<<

LECTURE NOTES ON DYNAMICS

Joseph M. Powers

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

updated 28 October 2019, 7:15pm 2

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

Preface 7

1 Introduction 9 1.1 Definitions...... 9 1.2 Motivatingexamples ...... 10 1.2.1 Re-entryflows...... 10 1.2.1.1 Bowshockwave ...... 10 1.2.1.2 Rarefaction(expansion)wave ...... 11 1.2.1.3 Momentumboundarylayer ...... 11 1.2.1.4 Thermalboundarylayer ...... 12 1.2.1.5 Vibrational relaxation effects ...... 12 1.2.1.6 Dissociation effects ...... 12 1.2.2 Rocketnozzleflows...... 12 1.2.3 Jetengineinlets...... 13

2 Governing equations 15 2.1 Mathematical preliminaries ...... 15 2.1.1 Vectorsandtensors...... 15 2.1.2 Gradient, divergence, and material derivatives ...... 16 2.1.3 Conservative andnon-conservative forms ...... 17 2.1.3.1 Conservativeform ...... 17 2.1.3.2 Non-conservativeform ...... 17 2.2 Summary of full set of compressible viscous equations ...... 20 2.3 Conservationaxioms ...... 21 2.3.1 Conservationofmass...... 22 2.3.1.1 Nonconservativeform ...... 22 2.3.1.2 Conservativeform ...... 22 2.3.1.3 Incompressible form ...... 22 2.3.2 Conservationoflinearmomenta ...... 22 2.3.2.1 Nonconservativeform ...... 22 2.3.2.2 Conservativeform ...... 23 2.3.3 Conservationofenergy ...... 24

3 4 CONTENTS

2.3.3.1 Nonconservativeform ...... 24 2.3.3.2 Mechanicalenergy ...... 25 2.3.3.3 Conservativeform ...... 25 2.3.3.4 Energyequationintermsofentropy ...... 26 2.3.4 Entropyinequality ...... 26 2.4 Constitutive relations ...... 29 2.4.1 Stress-strain rate relationship for Newtonian fluids ...... 29 2.4.2 Fourier’s law for heat conduction ...... 33 2.4.3 Variable first coefficient of , µ ...... 34 2.4.3.1 Typical values of µ forairandwater ...... 34 2.4.3.2 Common models for µ ...... 35 2.4.4 Variable second coefficient of viscosity, λ ...... 35 2.4.5 Variable thermal conductivity, k ...... 35 2.4.5.1 Typical values of k forairandwater ...... 35 2.4.5.2 Common models for k ...... 36 2.4.6 Thermalequationofstate ...... 36 2.4.6.1 Description ...... 36 2.4.6.2 Typical models ...... 36 2.4.7 Caloricequationofstate ...... 36 2.4.7.1 Description ...... 36 2.4.7.2 Typical models ...... 37 2.5 Specialcasesofgoverningequations ...... 37 2.5.1 One-dimensionalequations...... 37 2.5.2 Eulerequations ...... 38 2.5.3 Incompressible Navier-Stokes equations ...... 38

3 review 39 3.1 Preliminary mathematical concepts ...... 39 3.2 Summaryofthermodynamicconcepts...... 40 3.3 Maxwell relations and secondary properties ...... 44 3.3.1 Internal energy from thermal ...... 45 3.3.2 Soundspeedfromthermalequationofstate ...... 47 3.4 Canonicalequationsofstate ...... 53 3.5 Isentropicrelations ...... 55

4 One-dimensional compressible flow 61 4.1 Generalized one-dimensional equations ...... 62 4.1.1 Mass...... 63 4.1.2 Momentum ...... 63 4.1.3 Energy...... 65 4.1.4 Influencecoefficients ...... 72 4.2 Flowwithareachange ...... 73

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

4.2.1 Isentropic relations ...... 73 4.2.2 Sonicproperties...... 81 4.2.3 Effectofareachange ...... 82 4.2.4 Choking ...... 84 4.3 Normalshockwaves ...... 87 4.3.1 Governingequations ...... 88 4.3.2 Rayleigh line ...... 89 4.3.3 Hugoniotcurve ...... 89 4.3.4 Solution procedure for general equations of state ...... 91 4.3.5 Calorically perfect ideal gas solutions ...... 91 4.3.6 Acoustic limit ...... 102 4.3.7 Non-idealgassolutions ...... 103 4.4 Flowwithareachangeandnormalshocks ...... 107 4.4.1 Convergingnozzle...... 107 4.4.2 Converging-diverging nozzle ...... 108 4.5 Flow with –Fanno flow ...... 111 4.6 Flow with –Rayleigh flow ...... 117 4.7 Numerical solution of the shock tube problem ...... 122 4.7.1 One-steptechniques ...... 122 4.7.2 Lax-Friedrichs technique ...... 123 4.7.3 Lax-Wendrofftechnique ...... 123

5 Steady supersonic two-dimensional flow 125 5.1 Two-dimensionalequations...... 126 5.1.1 Conservativeform...... 126 5.1.2 Non-conservativeform ...... 126 5.2 Machwaves ...... 126 5.3 Obliqueshockwaves ...... 127 5.4 Smalldisturbancetheory...... 138 5.5 CenteredPrandtl-Meyerrarefaction ...... 141 5.6 Waveinteractionsandreflections ...... 145 5.6.1 Oblique shock reflected from a wall ...... 145 5.6.2 Oblique shock intersection ...... 146 5.6.3 Shockstrengthening ...... 146 5.6.4 Shockweakening ...... 147 5.7 Supersonic flow over airfoils ...... 147 5.7.1 Flatplateatangleofattack ...... 148 5.7.2 Diamond-shapedairfoil...... 152 5.7.3 Generalcurvedairfoil...... 153 5.7.4 Transonictransition ...... 153

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

6 Linear flow analysis 155 6.1 Formulation...... 155 6.2 Subsonicflow ...... 155 6.2.1 Prandtl-Glauretrule ...... 156 6.2.2 Flowoverwavywall ...... 156 6.3 Supersonicflow ...... 156 6.3.1 D’Alembert’s solution ...... 156 6.3.2 Flowoverwavywall ...... 156

7 Viscous flow 157 7.1 Governingequations ...... 157 7.2 Couetteflow...... 158 7.3 Suddenlyacceleratedflatplate...... 161 7.3.1 Formulation...... 161 7.3.2 Velocityprofile ...... 162 7.4 StartingtransientforplaneCouetteflow ...... 162 7.5 Blasiusboundarylayer ...... 163 7.5.1 Formulation...... 163 7.5.2 Wallshearstress ...... 163

8 Acoustics 165 8.1 Formulation...... 165 8.2 Planarwaves ...... 166 8.3 Sphericalwaves ...... 166

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

These are a set of class notes for a gas dynamics/viscous flow course taught to juniors in Aerospace Engineering at the University of Notre Dame during the mid 1990s. The course builds upon foundations laid in an earlier course where the emphasis was on subsonic ideal flows. Consequently, it is expected that the student has some familiarity with many concepts such as material derivatives, control volume analysis, derivation of governing equations, etc. Additionally, first courses in thermodynamics and differential equations are probably necessary. Even a casual reader will find gaps, errors, and inconsistencies. The author welcomes comments and corrections. It is also noted that these notes have been influenced by a variety of standard references, which are sporadically and incompletely noted in the text. Some of the key references which were important in the development of these notes are the texts of Shapiro, Liepmann and Roshko, Anderson, Courant and Friedrichs, Hughes and Brighton, White, Sonntag and Van Wylen, and Zucrow and Hoffman.

At this stage, if anyone outside Notre Dame finds these useful, they are free to make copies. Full information on the course is found at http://www.nd.edu/ powers/ame.30332. ∼ Joseph M. Powers [email protected] http://www.nd.edu/ powers Notre Dame, Indiana;∼ USA CC BY: $\ = 28 October 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 October 2019, J. M. Powers. Chapter 1

Introduction

Suggested Reading:

Anderson, Chapter 1: pp. 1-31

1.1 Definitions

The topic of this course is the of compressible and viscous flow.

Where does aerodynamics rest in the taxonomy of mechanics?

Aerodynamics–a branch of dynamics that deals with the motion of air and other gaseous fluids and with the forces acting on bodies in motion relative to such fluids (e.g. airplanes)

We can say that aerodynamics is a subset of ( ) ⊂ fluid dynamics since air is but one type of fluid, • ⊂ fluid mechanics since dynamics is part of mechanics, • ⊂ mechanics since fluid mechanics is one class of mechanics. •

Mechanics–a branch of physical science that deals with forces and the motion of bodies traditionally broken into:

kinematics–study of motion without regard to causality • dynamics (kinetics)–study of forces which give rise to motion • Examples of other subsets of mechanics:

9 10 CHAPTER 1. INTRODUCTION

solid mechanics • quantum mechanics • celestial mechanics • relativistic mechanics • quantum-electrodynamics (QED) • magneto-hydrodynamics (MHD) • Recall the definition of a fluid:

Fluid–a material which moves when a shear force is applied.

Recall that solids can, after a small displacement, relax to an equilibrium configuration when a shear force is applied.

Recall also that both liquids and are fluids

The motion of both liquids and gases can be affected by and shear forces. While shear forces are important for both types of fluids, the influence of compressibility in gases is generally more significant.

The thrust of this class will be to understand how to model the effects of compressibility and shear forces and how this impacts the design of aerospace vehicles.

1.2 Motivating examples

The following two examples serve to illustrate why knowledge of compressibility and shear effects is critical.

1.2.1 Re-entry flows A range of phenomena are present in the re-entry of a vehicle into the atmosphere. This is an example of an external flow. See Figure 1.1.

1.2.1.1 Bow suddenly raises density, and pressure of shocked air; consider normal shock • in ideal air

– ρ =1.16 kg/m3 ρ =6.64 kg/m3 (over five times as dense!!) o → s – T = 300 K T =6, 100 K (hot as the sun’s surface !!) o → s

CC BY-NC-ND. 28 October 2019, J. M. Powers. 1.2. MOTIVATING EXAMPLES 11

far-field acoustic wave

rarefaction waves

viscous and thermal boundary layers

Oblique Shock Wave Ambient Air

Normal Shock Wave

Figure 1.1: phenomena in re-entry

– P =1.0 atm P = 116.5 atm (tremendous force change!!) o → s – sudden transfer of energy from kinetic (ordered) to thermal (random)

introduces inviscid /vorticity layer into post-shocked flow • normal shock standing off leading edge • conical oblique shock away from leading edge • acoustic wave in far field •

1.2.1.2 Rarefaction (expansion) wave lowers density, temperature, and pressure of air continuously and significantly • interactions with bow shock weaken bow shock •

1.2.1.3 Momentum occurs in thin layer near surface where velocity relaxes from freestream to zero to • satisfy the no-slip condition

necessary to predict viscous drag forces on body •

CC BY-NC-ND. 28 October 2019, J. M. Powers. 12 CHAPTER 1. INTRODUCTION

1.2.1.4 Thermal boundary layer as fluid decelerates in momentum boundary layer kinetic energy is converted to thermal • energy temperature rises can be significant (> 1, 000 K) • 1.2.1.5 Vibrational relaxation effects energy partitioned into vibrational modes in addition to translational • lowers temperature that would otherwise be realized • important for air above 800 K • unimportant for monatomic gases • 1.2.1.6 Dissociation effects effect which happens when multi-atomic molecules split into constituent atoms • O totally dissociated into O near 4, 000 K • 2 N totally dissociated into N near 9, 000 K • 2 For T > 9, 000 K, ionized plasmas begin to form • Vibrational relaxation, dissociation, and ionization can be accounted for to some extent by introducing a temperature-dependent specific heat cv(T )

1.2.2 Rocket nozzle flows The same essential ingredients are present in flows through rocket nozzles. This is an example of an internal flow, see Figure 1.2

viscous and thermal boundary layers

burning solid rocket fuel

burning solid rocket fuel

possible normal shock

Figure 1.2: Fluid mechanics phenomena in rocket nozzles

Some features:

CC BY-NC-ND. 28 October 2019, J. M. Powers. 1.2. MOTIVATING EXAMPLES 13

well-modelled as one-dimensional flow • large thrust relies on subsonic to supersonic transition in a converging-diverging nozzle • away from design conditions normal shocks can exist in nozzle • viscous and thermal boundary layers must be accounted for in design • 1.2.3 Jet engine inlets The same applies for the internal flow inside a jet engine, see Figure 1.3

viscous and thermal boundary layers oblique shock

compressor turbine combustor exhaust inlet

Figure 1.3: Fluid mechanics phenomena in jet engine inlet

CC BY-NC-ND. 28 October 2019, J. M. Powers. 14 CHAPTER 1. INTRODUCTION

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

Governing equations

Suggested Reading:

Hughes and Brighton, Chapter 3: pp. 44-64 Liepmann and Roshko, Chapter 7: pp. 178-190, Chapter 13: pp. 305-313, 332-338 Anderson, Chapter 2: pp. 32-44; Chapter 6: pp. 186-205

The equations which govern a wide variety of these flows are the compressible Navier- Stokes equations. In general they are quite complicated and require numerical solution. We will only consider small subsets of these equations in practice, but it is instructive to see them in full glory at the outset.

2.1 Mathematical preliminaries

A few concepts which may be new or need re-emphasis are introduced here.

2.1.1 Vectors and tensors One way to think of vectors and tensors is as follows:

first order tensor: vector, associates a scalar with any direction in space, column • matrix

second order tensor: tensor-associates a vector with any direction in space, two- • dimensional matrix

third order tensor-associates a second order tensor with any direction in space, three- • dimensional matrix

fourth order tensor-... • 15 16 CHAPTER 2. GOVERNING EQUATIONS

Here a vector, denoted by boldface, denotes a quantity which can be decomposed as a sum of scalars multiplying orthogonal basis vectors, i.e.:

v = ui + vj + wk (2.1)

2.1.2 Gradient, divergence, and material derivatives The “del” operator, , is as follows: ∇ ∂ ∂ ∂ i + j + k (2.2) ∇ ≡ ∂x ∂y ∂z Recall the definition of the material derivative also known as the substantial or total derivative:

d ∂ + v (2.3) dt ≡ ∂t ·∇ where

Example 2.1 Does v = v = v? · ∇ ∇ · ∇ ∂ ∂ ∂ v = u + v + w (2.4) · ∇ ∂x ∂y ∂z

∂u ∂v ∂w v = + + (2.5) ∇ · ∂x ∂y ∂z

∂u ∂v ∂w ∂x ∂x ∂x ∂u ∂v ∂w v = ∂y ∂y ∂y (2.6) ∇  ∂u ∂v ∂w  ∂z ∂z ∂z   So, no.

Here the quantity v is an example of a second order tensor. Also ∇ v v div (2.7) ·∇≡ v div v (2.8) ∇ · ≡ v grad v (2.9) ∇ ≡ φ grad φ (2.10) ∇ ≡

CC BY-NC-ND. 28 October 2019, J. M. Powers. 2.1. MATHEMATICAL PRELIMINARIES 17

2.1.3 Conservative and non-conservative forms

T If hi is a column vector of N variables, e.g. hi = [h1, h2, h3, ...hN ] , and fi(hi) gi(hi) are a column vectors of N functions of the variables hi, and all variables are functions of x and t, hi = hi(x, t), fi(hi(x, t)), gi(hi(x, t)) then a system of partial differential equations is in conservative form iff the system can be written as follows:

∂ ∂ h + (f (h )) = g (h ) (2.11) ∂t i ∂x i i i i A system not in this form is in non-conservative form

2.1.3.1 Conservative form Advantages

naturally arises from control volume derivation of governing equations • clearly exposes groups of terms which are conserved • easily integrated in certain special cases • most natural form for deriving normal shock jump equations • the method of choice for numerical simulations • Disadvantages

lengthy • not commonly used • difficult to see how individual variables change • 2.1.3.2 Non-conservative form Advantages

compact • commonly used • can see how individual variables change • Disadvantages

often difficult to use to get solutions to problems •

CC BY-NC-ND. 28 October 2019, J. M. Powers. 18 CHAPTER 2. GOVERNING EQUATIONS

gives rise to artificial instabilities if used in numerical simulation •

Example 2.2 Kinematic wave equation

The kinematic wave equation in non-conservative form is

∂u ∂u + u =0 (2.12) ∂t ∂x This equation has the same mathematical form as inviscid equations of gas dynamics which give rise to discontinuous shock waves. Thus understanding the solution of this simple equation is very useful in understanding equations with more physical significance.

∂u ∂ u2 Since u ∂x = ∂x 2 the kinematic wave equation in conservative form is as follows:   ∂u ∂ u2 + =0 (2.13) ∂t ∂x 2   u2 Here hi = u,fi = 2 ,gi = 0.

∂ Consider the special case of a steady state ∂t 0. Then the conservative form of the equation can be integrated! ≡

d u2 =0 (2.14) dx 2   u2 u2 = o (2.15) 2 2 u = u (2.16) ± o

Now u = u satisfies the equation and so does u = u . These are both smooth solutions. In o − o addition, combinations also satisfy, e.g. u = uo,x< 0; u = uo, x 0. This is a discontinuous solution. − ≥∂u Also note the solution is not unique. This is a consequence of the u ∂x non-linearity. This is an example of a type of shock wave. Which solution is achieved generally depends on terms we have neglected, especially unsteady terms.

Example 2.3 Burger’s equation

Burger’s equation in non-conservative form is

∂u ∂u ∂2u + u = ν (2.17) ∂t ∂x ∂x2

CC BY-NC-ND. 28 October 2019, J. M. Powers. 2.1. MATHEMATICAL PRELIMINARIES 19

This equation has the same mathematical form as viscous equations of gas dynamics which give rise to spatially smeared shock waves.

Place this in conservative form: ∂u ∂u ∂ ∂u + u ν = 0 (2.18) ∂t ∂x − ∂x ∂x ∂u ∂ u2 ∂ ∂u + ν = 0 (2.19) ∂t ∂x 2 − ∂x ∂x     ∂u ∂ u2 ∂u + ν = 0 (2.20) ∂t ∂x 2 − ∂x   ∂ Here, this equation is not strictly in conservative form as it still involves derivatives inside the ∂x operator.

∂ Consider the special case of a steady state ∂t 0. Then the conservative form of the equation can be integrated! ≡ d u2 du ν =0 (2.21) dx 2 − dx   Let u u as x (consequently ∂u 0 as x ) and u(0)= 0 so → o → −∞ ∂x → → −∞ u2 du u2 ν = o (2.22) 2 − dx 2 du 1 ν = u2 u2 (2.23) dx 2 − o du dx = (2.24) u2 u2 2ν − o du dx = (2.25) u2 u2 − 2ν Z o − Z 1 − u x tanh 1 = + C (2.26) uo uo −2ν u u(x)= u tanh o x + Cu (2.27) o −2ν o u(0)=0= uo tanh (Cuo) C = 0 (2.28) u u(x)= u tanh o x (2.29) o −2ν lim u(x)= uo (2.30) x→−∞

lim u(x)= uo (2.31) x→∞ − Note same behavior in far field as kinematic wave equation • 2ν continuous adjustment from uo to uo in a zone of thickness • − uo zone thickness 0 as ν 0 • → → inviscid shock is limiting case of viscously resolved shock • Figure 2.1 gives a plot of the solution to both the kinematic wave equation and Burger’s equation.

CC BY-NC-ND. 28 October 2019, J. M. Powers. 20 CHAPTER 2. GOVERNING EQUATIONS

u u

uo uo

x x

-uo -uo

Kinematic Wave Equation Solution Burger's Equation Solution Discontinuous Shock Wave Smeared Shock Wave Shock Thickness ~ 2 ν / uo

Figure 2.1: Solutions to the kinematic wave equation and Burger’s equation

2.2 Summary of full set of compressible viscous equa- tions

A complete set of equations is given below. These are the compressible Navier-Stokes equa- tions for an isotropic Newtonian fluid with variable properties dρ + ρ v = 0 [1] (2.32) dt ∇· dv ρ = P + τ + ρg [3] (2.33) dt −∇ ∇· de ρ = q P v + τ : v [1] (2.34) dt −∇ · − ∇· ∇ τ = µ v + vT + λ ( v) I [6] (2.35) ∇ ∇ ∇· q = k T [3] (2.36)  − ∇ µ = µ (ρ, T ) [1] (2.37) λ = λ (ρ, T ) [1] (2.38) k = k (ρ, T ) [1] (2.39) P = P (ρ, T ) [1] (2.40) e = e (ρ, T ) [1] (2.41)

The numbers in brackets indicate the number of equations. Here the unknowns are ρ–density kg/m3 (scalar-1 variable) • v–velocity m/s (vector- 3 variables) • P –pressure N/m2 (scalar- 1 variable) • e–internal energy J/kg (scalar- 1 variable) •

CC BY-NC-ND. 28 October 2019, J. M. Powers. 2.3. CONSERVATION AXIOMS 21

T –temperature K (scalar - 1 variable) • τ –viscous stress N/m2 (symmetric tensor - 6 variables) • q–heat flux vector–W/m2 (vector - 3 variables) • µ–first coefficient of viscosity Ns/m2 (scalar - 1 variable) • λ–second coefficient of viscosity Ns/m2 (scalar - 1 variable) • k–thermal conductivity W/(m2K) (scalar - 1 variable) • Here g is the constant gravitational acceleration and I is the identity matrix. Total–19 variables Points of the exercise 19 equations; 19 unknowns • conservation axioms–postulates (first three equations) • constitutive relations–material dependent (remaining equations) • review of vector notation and operations • Exercise: Determine the three Cartesian components of τ for a) a compressible Newtonian fluid, and b) an incompressible Newtonian fluid, in which∇· v = 0. ∇· This system of equations must be consistent with the second law of thermodynamics. Defining the entropy s by the Gibbs relation: 1 Tds = de + Pd (2.42) ρ   ds de d 1 T = + P (2.43) dt dt dt ρ   the second law states: ds q ρ (2.44) dt ≥ −∇· T In practice, this places some simple restrictions on the constitutive relations. It will be sometimes useful to write this in terms of the specific volume, v 1/ρ. This can be confused with the y component of velocity but should be clear in context.≡

2.3 Conservation axioms

Conservation principles are axioms of mechanics and represent statements that cannot be proved. In that they provide predictions which are consistent with empirical observations, they are useful.

CC BY-NC-ND. 28 October 2019, J. M. Powers. 22 CHAPTER 2. GOVERNING EQUATIONS

2.3.1 Conservation of mass This principle states that in a material volume (a volume which always encompasses the same fluid particles), the mass is constant.

2.3.1.1 Nonconservative form dρ + ρ v = 0 (2.45) dt ∇· This can be expanded using the definition of the material derivative to form

∂ρ ∂ρ ∂ρ ∂ρ ∂u ∂v ∂w + u + v + w + ρ + + = 0 (2.46) ∂t ∂x ∂y ∂x ∂x ∂y ∂z   2.3.1.2 Conservative form Using the product rule gives

∂ρ ∂(ρu) ∂(ρv) ∂(ρw) + + + = 0 (2.47) ∂t ∂x ∂y ∂z The equation essentially says that the net accumulation of mass within a control volume is attributable to the net flux of mass in and out of the control volume. In Gibbs notation this is ∂ρ + (ρv) = 0 (2.48) ∂t ∇·

2.3.1.3 Incompressible form Iff the fluid is defined to be incompressible, dρ/dt 0, the consequence is ≡ v =0, or (2.49) ∇· ∂u ∂v ∂w + + = 0 (2.50) ∂x ∂y ∂z As this course is mainly concerned with compressible flow, this will not be often used.

2.3.2 Conservation of linear momenta This is really Newton’s Second Law of Motion ma = F P 2.3.2.1 Nonconservative form dv ρ = P + τ + ρg (2.51) dt −∇ ∇· ρ: mass/volume •

CC BY-NC-ND. 28 October 2019, J. M. Powers. 2.3. CONSERVATION AXIOMS 23

dv : acceleration • dt P, τ : surface forces/volume • −∇ ∇· ρg: body force/volume •

Example 2.4 Expand the term τ ∇ · T ∂ τ + ∂ τ + ∂ τ τxx τxy τxz ∂x xx ∂y yx ∂z zx τ ∂ ∂ ∂ ∂ τ + ∂ τ + ∂ τ = ∂x ∂y ∂z τyx τyy τyz =  ∂x xy ∂y yy ∂z zy  (2.52) ∇ ·   ∂ ∂ ∂ τzx τzy τzz τxz + τyz + τzz   ∂x ∂y ∂z     

This is a vector equation as there are three components of momenta. Let’s consider the x momentum equation for example. du ∂P ∂τ ∂τ ∂τ ρ = + xx + yx + zx + ρg (2.53) dt − ∂x ∂x ∂y ∂z x Now expand the material derivative: ∂u ∂u ∂u ∂u ∂P ∂τ ∂τ ∂τ ρ + ρu + ρv + ρw = + xx + yx + zx + ρg (2.54) ∂t ∂x ∂y ∂z − ∂x ∂x ∂y ∂z x Equivalent equations exist for y and z linear momentum: ∂v ∂v ∂v ∂v ∂P ∂τ ∂τ ∂τ ρ + ρu + ρv + ρw = + xy + yy + zy + ρg (2.55) ∂t ∂x ∂y ∂z − ∂y ∂x ∂y ∂z y

∂w ∂w ∂w ∂w ∂P ∂τ ∂τ ∂τ ρ + ρu + ρv + ρw = + xz + yz + zz + ρg (2.56) ∂t ∂x ∂y ∂z − ∂z ∂x ∂y ∂z z

2.3.2.2 Conservative form Multiply the mass conservation principle by u so that it has the same units as the momentum equation and add to the x momentum equation:

∂ρ ∂(ρu) ∂(ρv) ∂(ρw) u + u + u + u = 0 (2.57) ∂t ∂x ∂y ∂z

∂u ∂u ∂u ∂u ∂P ∂τ ∂τ ∂τ +ρ + ρu + ρv + ρw = + xx + yx + zx + ρg (2.58) ∂t ∂x ∂y ∂z − ∂x ∂x ∂y ∂z x

CC BY-NC-ND. 28 October 2019, J. M. Powers. 24 CHAPTER 2. GOVERNING EQUATIONS

Using the product rule, this yields: ∂ (ρu) ∂ (ρuu) ∂ (ρvu) ∂ (ρwu) ∂P ∂τ ∂τ ∂τ + + + = + xx + yx + zx + ρg (2.59) ∂t ∂x ∂y ∂z − ∂x ∂x ∂y ∂z x The extension to y and z momenta is straightforward: ∂ (ρv) ∂ (ρuv) ∂ (ρvv) ∂ (ρwv) ∂P ∂τ ∂τ ∂τ + + + = + xy + yy + zy + ρg (2.60) ∂t ∂x ∂y ∂z − ∂y ∂x ∂y ∂z y ∂ (ρw) ∂ (ρuw) ∂ (ρvw) ∂ (ρww) ∂P ∂τ ∂τ ∂τ + + + = + xz + yz + zz + ρg (2.61) ∂t ∂x ∂y ∂z − ∂z ∂x ∂y ∂z z In vector form this is written as follows: ∂ (ρv) + (ρvv)= P + τ + ρg (2.62) ∂t ∇· −∇ ∇· As with the mass equation, the time derivative can be interpreted as the accumulation of linear momenta within a control volume and the divergence term can be interpreted as the flux of linear momenta into the control volume. The accumulation and flux terms are balanced by forces, both surface and body.

2.3.3 This principle really is the first law of thermodynamics, which states the change in internal energy of a body is equal to the heat added to the body minus the work done by the body; Eˆ Eˆ = Q W (2.63) 2 − 1 12 − 12 The Eˆ here includes both internal energy and kinetic energy and is written for an extensive system: 1 Eˆ = ρV e + v v (2.64) 2 ·   2.3.3.1 Nonconservative form The equation we started with (which is in non-conservative form) de ρ = q P v + τ : v (2.65) dt −∇ · − ∇· ∇ is simply a careful expression of the simple idea de = dq dw with attention paid to sign − conventions, etc. ρ de : change in internal energy /volume • dt q: net heat transfer into fluid/volume • −∇· P v: net work done by fluid due to pressure force/volume (force deformation) • ∇· × τ : v: net work done by fluid due to viscous force/volume (force deformation) • − ∇ ×

CC BY-NC-ND. 28 October 2019, J. M. Powers. 2.3. CONSERVATION AXIOMS 25

2.3.3.2 Mechanical energy Taking the dot product of the velocity v with the linear momentum principle yields the mechanical energy equation (here expressed in conservative form): ∂ 1 1 ρ (v v) + ρv (v v) = v P + v ( τ )+ ρv g (2.66) ∂t 2 · ∇· 2 · − ·∇ · ∇· ·     This can be interpreted as saying the kinetic energy (or mechanical energy) changes due to motion in the direction of a force imbalance • – v P − ·∇ – v ( τ ) · ∇· motion in the direction of a body force • Exercise: Add the product of the mass equation and u2/2 to the product of u and the one dimensional linear momentum equation: ∂u ∂u ∂P ∂τ u ρ + ρu = u + xx + ρg (2.67) ∂t ∂x − ∂x ∂x x     to form the conservative form of the one-dimensional mechanical energy equation: ∂ 1 ∂ 1 ∂P ∂τ ρu2 + ρu3 = u + u xx + ρug (2.68) ∂t 2 ∂x 2 − ∂x ∂x x     2.3.3.3 Conservative form When we multiply the mass equation by e, we get ∂ρ ∂(ρu) ∂(ρv) ∂(ρw) e + e + e + e = 0 (2.69) ∂t ∂x ∂y ∂z Adding this to the nonconservative energy equation gives ∂ (ρe)+ (ρve)= q P v + τ : v (2.70) ∂t ∇· −∇ · − ∇· ∇ Adding to this the mechanical energy equation gives the conservative form of the energy equation: ∂ 1 1 ρ e + v v + ρv e + v v = q (P v)+ (τ v)+ρv g (2.71) ∂t 2 · ∇· 2 · −∇· −∇· ∇· · ·       which is often written as ∂ 1 1 P ρ e + v v + ρv e + v v + = q + (τ v)+ ρv g (2.72) ∂t 2 · ∇· 2 · ρ −∇ · ∇· · ·      

CC BY-NC-ND. 28 October 2019, J. M. Powers. 26 CHAPTER 2. GOVERNING EQUATIONS

2.3.3.4 Energy equation in terms of entropy Recall the Gibbs relation which defines entropy s:

ds de d 1 de P dρ T = + P = (2.73) dt dt dt ρ dt − ρ2 dt   so de ds P dρ ρ = ρT + (2.74) dt dt ρ dt also from the conservation of mass 1 dρ v = (2.75) ∇· −ρ dt Substitute into nonconservative energy equation: ds P dρ P dρ ρT + = q + + τ : v (2.76) dt ρ dt −∇ · ρ dt ∇ Solve for entropy change: ds 1 1 ρ = q + τ : v (2.77) dt −T ∇· T ∇ Two effects change entropy:

heat transfer • viscous work • Note the work of the pressure force does not change entropy; it is reversible work.

If there are no viscous and heat transfer effects, there is no mechanism for entropy change; ds/dt = 0; the flow is isentropic.

2.3.4 Entropy inequality The first law can be used to reduce the second law to a very simple form. Starting with q 1 q = q T (2.78) ∇· T T ∇· − T 2 ·∇   so 1 q q q = T (2.79) −T ∇· −∇ · T − T 2 ·∇ Substitute into the first law:   ds q q 1 ρ = T + τ : v (2.80) dt −∇ · T − T 2 ·∇ T ∇   CC BY-NC-ND. 28 October 2019, J. M. Powers. 2.3. CONSERVATION AXIOMS 27

Recall the second law of thermodynamics: ds q ρ (2.81) dt ≥ −∇· T   Substituting the first law into the second law thus yields: q 1 T + τ : v 0 (2.82) −T 2 ·∇ T ∇ ≥ Our constitutive theory for q and τ must be constructed to be constructed so as not to violate the second law.

CC BY-NC-ND. 28 October 2019, J. M. Powers. 28 CHAPTER 2. GOVERNING EQUATIONS

Exercise: Beginning with the unsteady, two-dimensional, compressible Navier-Stokes equations with no body force in conservative form (below), show all steps necessary to reduce these to the following non-conservative form.

Conservative form

∂ρ ∂ ∂ + (ρu)+ (ρv)=0 ∂t ∂x ∂y ∂ ∂ ∂ (ρu)+ (ρuu + P τ )+ (ρuv τ )=0 ∂t ∂x − xx ∂y − yx ∂ ∂ ∂ (ρv)+ (ρvu τ )+ (ρvv + P τ )=0 ∂t ∂x − xy ∂y − yy ∂ 1 ρ e + u2 + v2 ∂t 2    ∂ 1 P  + ρu e + u2 + v2 + (uτ + vτ )+ q ∂x 2 ρ − xx xy x     ∂ 1 P + ρv e + u2 + v2 + (uτ + vτ )+ q =0 ∂y 2 ρ − yx yy y      Non-conservative form

∂ρ ∂ρ ∂ρ ∂u ∂v + u + v + ρ + =0 ∂t ∂x ∂y ∂x ∂y     ∂u ∂u ∂u ∂P ∂τ ∂τ ρ + u + v = + xx + yx ∂t ∂x ∂y − ∂x ∂x ∂y   ∂v ∂v ∂v ∂P ∂τ ∂τ ρ + u + v = + xy + yy ∂t ∂x ∂y − ∂y ∂x ∂y   ∂e ∂e ∂e ρ + u + v ∂t ∂x ∂y   ∂q ∂q = x + y − ∂x ∂y   ∂u ∂v P + − ∂x ∂y   ∂u ∂v ∂u ∂v +τ + τ + τ + τ xx ∂x xy ∂x yx ∂y yy ∂y

CC BY-NC-ND. 28 October 2019, J. M. Powers. 2.4. CONSTITUTIVE RELATIONS 29

2.4 Constitutive relations

These are determined from experiments and provide sometimes good and sometimes crude models for microstructurally based phenomena.

2.4.1 Stress-strain rate relationship for Newtonian fluids Perform the experiment described in Figure 2.2.

Force = F Velocity = U h

Figure 2.2: Schematic of experiment to determine stress-strain-rate relationship

The following results are obtained, Figure 2.3:

h1 A4 F h F 2 A3

h3 A2

h4 A1

U U A4 > A3 > A2 > A1 h4 > h3 > h2 > h1

Figure 2.3: Force (N) vs. velocity (m/s)

Note for constant plate velocity U

small gap width h gives large force F • large cross-sectional area A gives large force F • When scaled by h and A, for a single fluid, the curve collapses to a single curve, Figure 2.4:

The viscosity is defined as the ratio of the applied stress τyx = F/A to the strain rate ∂u ∂y .

CC BY-NC-ND. 28 October 2019, J. M. Powers. 30 CHAPTER 2. GOVERNING EQUATIONS

F/A

µ

1

U/h

Figure 2.4: Stress (N/m2) vs. strain rate (1/s)

τyx µ ∂u (2.83) ≡ ∂y Here the first subscript indicates the face on which the force is acting, here the y face. The second subscript indicates the direction in which the force takes, here the x direction. In general viscous stress is a tensor quantity. In full detail it is as follows:

∂u ∂u ∂u ∂v ∂u ∂w ∂x + ∂x ∂y + ∂x ∂z + ∂x τ ∂v ∂u ∂v ∂v ∂v ∂w = µ ∂x + ∂y ∂y + ∂y ∂z + ∂y  ∂w ∂u ∂w ∂v ∂w ∂w  ∂x + ∂z ∂y + ∂z ∂z + ∂z   ∂u ∂v ∂w ∂x + ∂y + ∂z 0 0 ∂u ∂v ∂w +λ 0 ∂x + ∂y + ∂z 0 (2.84)  ∂u ∂v ∂w  0 0 ∂x + ∂y + ∂z   This is simply an expanded form of that written originally:

τ = µ v + vT + λ ( v) I (2.85) ∇ ∇ ∇· Here λ is the second coefficient of viscosity. It is irrelevant in incompressible flows and notoriously difficult to measure in compressible flows. It has been the source of controversy for over 150 years. Commonly, and only for convenience, people take Stokes’ Assumption: 2 λ µ (2.86) ≡ −3 It can be shown that this results in the mean mechanical stress being equivalent to the thermodynamic pressure.

CC BY-NC-ND. 28 October 2019, J. M. Powers. 2.4. CONSTITUTIVE RELATIONS 31

It can also be shown that the second law is satisfied if 2 µ 0 and λ µ (2.87) ≥ ≥ −3

Example 2.5 Couette Flow

Use the linear momentum principle and the constitutive theory to show the velocity profile between two plates is linear. The lower plate at y = 0 is stationary; the upper plate at y = h is moving at velocity U. Assume v = u(y)i +0j +0k. Assume there is no imposed pressure gradient or body force. Assume constant viscosity µ. Since u = u(y), v =0, w = 0, there is no fluid acceleration. ∂u ∂u ∂u ∂u + u + v + w =0+0+0+0=0 (2.88) ∂t ∂x ∂y ∂z Since no pressure gradient or body force the linear momentum principle is simply ∂τ 0= yx (2.89) ∂y With the Newtonian fluid ∂ ∂u 0= µ (2.90) ∂y ∂y   With constant µ and u = u(y) we have:

d2u µ =0 (2.91) dx2 Integrating we find u = Ay + B (2.92) Use the boundary conditions at y = 0 and y = h to give A and B: U A =0, B = (2.93) h so U u(y)= y (2.94) h

Example 2.6 Poiseuille Flow

Consider flow between a slot separated by two plates, the lower at y = 0, the upper at y = h, both plates stationary. The flow is driven by a pressure difference. At x = 0, P = Po; at x = L, P = P1. The fluid has constant viscosity µ. Assuming the flow is steady, there is no body force, pressure varies only with x, and that the velocity is only in the x direction and only a function of y; i.e. v = u(y) i, find the velocity profile u(y) parameterized by Po, P1, h, and µ.

CC BY-NC-ND. 28 October 2019, J. M. Powers. 32 CHAPTER 2. GOVERNING EQUATIONS

As before there is no acceleration and the x momentum equation reduces to:

∂P ∂2u 0= + µ (2.95) − ∂x ∂y2

First let’s find the pressure field; take ∂/∂x:

∂2P ∂ ∂2u 0= + µ (2.96) − ∂x2 ∂x ∂y2   ∂2P ∂2 ∂u changing order of differentiation: 0 = + µ (2.97) − ∂x2 ∂y2 ∂x   ∂2P d2P 0= = (2.98) − ∂x2 − dx2 dP = A (2.99) dx P = Ax + B (2.100)

apply boundary conditions : P (0) = Po P (L)= P1 (2.101) x P (x)= P + (P P ) (2.102) o 1 − o L dP (P P ) so = 1 − o (2.103) dx L (P P ) d2u substitute into momentum: 0 = 1 − o + µ (2.104) − L dy2 d2u (P P ) = 1 − o (2.105) dy2 µL du (P P ) = 1 − o y + C (2.106) dy µL 1 (P P ) u(y)= 1 − o y2 + C y + C (2.107) 2µL 1 2

boundary conditions: u(0)=0= C2 (2.108) (P P ) u(h)=0= 1 − o h2 + C h + 0 (2.109) 2µL 1 (P P ) C = 1 − o h (2.110) 1 − 2µL (P P ) u(y)= 1 − o y2 yh (2.111) 2µL − du (P P )  wall shear: = 1 − o (2y h) (2.112) dy 2µL − du (P P ) τ = µ = h 1 − o (2.113) wall dy − 2L y=0

Exercise: Consider flow between a slot separated by two plates, the lower at y = 0, the upper at y = h, with the bottom plate stationary and the upper plate moving at velocity

CC BY-NC-ND. 28 October 2019, J. M. Powers. 2.4. CONSTITUTIVE RELATIONS 33

U. The flow is driven by a pressure difference and the motion of the upper plate. At x = 0, P = Po; at x = L, P = P1. The fluid has constant viscosity µ. Assuming the flow is steady, there is no body force, pressure varies only with x, and that the velocity is only in the x direction and only a function of y; i.e. v = u(y)i, a) find the velocity profile u(y) parameterized by Po, P1, h, U and µ; b) Find U such that there is no net mass flux between the plates.

2.4.2 Fourier’s law for heat conduction It is observed in experiment that heat moves from regions of high temperature to low tem- perature Perform the experiment described in Figure 2.5.

x T A q To

L

T > To

Figure 2.5: Schematic of experiment to determine thermal conductivity

The following results are obtained, Figure 2.6:

Q Q Q A L1 t3 3 L t2 A2 2

t1 A1 L1

T T T A > A > A L3 > L 2 > L1 t3 > t2 > t1 3 2 1

Figure 2.6: Heat transferred (J) vs. temperature (K)

Note for constant temperature of the high temperature reservoir T large time of heat transfer t gives large heat transfer Q • large cross-sectional area A gives large heat transfer Q • small length L gives large heat transfer Q • When scaled by L, t, and A, for a single fluid, the curve collapses to a single curve, Figure 2.7:

CC BY-NC-ND. 28 October 2019, J. M. Powers. 34 CHAPTER 2. GOVERNING EQUATIONS

Q/(A t)

k 1

T/L

Figure 2.7: heat flux vs. temperature gradient

The thermal conductivity is defined as the ratio of the flux of heat transfer q Q/(At) x ∼ to the temperature gradient ∂T T/L. − ∂x ∼ q k x (2.114) ≡ ∂T − ∂x so ∂T q = k (2.115) x − ∂x or in vector notation: q = k T (2.116) − ∇ Note with this form, the contribution from heat transfer to the entropy production is guaranteed positive if k 0. ≥ T T 1 k ∇ ·∇ + τ : v 0 (2.117) T 2 T ∇ ≥

2.4.3 Variable first coefficient of viscosity, µ In general the first coefficient of viscosity µ is a thermodynamic property which is a strong function of temperature and a weak function of pressure.

2.4.3.1 Typical values of µ for air and water

6 2 air at 300 K, 1 atm : 18.46 10− (Ns)/m • × 6 2 air at 400 K, 1 atm : 23.01 10− (Ns)/m • × 6 2 liquid water at 300 K, 1 atm : 855 10− (Ns)/m • ×

CC BY-NC-ND. 28 October 2019, J. M. Powers. 2.4. CONSTITUTIVE RELATIONS 35

6 2 liquid water at 400 K, 1 atm : 217 10− (Ns)/m • × Note viscosity of air an order of magnitude less than water • ∂µ > 0 for air, and gases in general • ∂T ∂µ < 0 for water, and liquids in general • ∂T 2.4.3.2 Common models for µ constant property: µ = µ • o T kinetic theory estimate for high temperature gas: µ (T )= µo • To q empirical data • 2.4.4 Variable second coefficient of viscosity, λ Very little data for any material exists for the second coefficient of viscosity. It only plays a role in compressible viscous flows, which are typically very high speed. Some estimates: Stokes’ hypothesis: λ = 2 µ, may be correct for monatomic gases • − 3 may be inferred from attenuation rates of sound waves • perhaps may be inferred from shock wave thicknesses • 2.4.5 Variable thermal conductivity, k In general thermal conductivity k is a thermodynamic property which is a strong function of temperature and a weak function of pressure.

2.4.5.1 Typical values of k for air and water

3 air at 300 K, 1 atm : 26.3 10− W/(mK) • × 3 air at 400 K, 1 atm : 33.8 10− W/(mK) • × 3 liquid water at 300 K, 1 atm : 613 10− W/(mK) • × 3 liquid water at 400 K, 1 atm : 688 10− W/(mK) (the liquid here is supersaturated) • × Note conductivity of air is one order of magnitude less than water • ∂k > 0 for air, and gases in general • ∂T ∂k > 0 for water in this range, generalization difficult • ∂T

CC BY-NC-ND. 28 October 2019, J. M. Powers. 36 CHAPTER 2. GOVERNING EQUATIONS

2.4.5.2 Common models for k constant property: k = k • o T kinetic theory estimate for high temperature gas: k (T )= ko • To q empirical data •

Exercise: Consider one-dimensional steady heat conduction in a fluid at rest. At x = 2 0 m at constant heat flux is applied qx = 10 W/m . At x = 1 m, the temperature is held constant at 300 K. Find T (y), T (0) and qx(1) for

3 liquid water with k = 613 10− W/(mK) • × 3 air with k = 26.3 10− W/(mK) • × 3 T air with k = 26.3 10− W/(mK) • × 300  q  2.4.6 Thermal equation of state 2.4.6.1 Description determined in static experiments • gives P as a function of ρ and T • 2.4.6.2 Typical models ideal gas: P = ρRT • first virial: P = ρRT (1 + b ρ) • 1 general virial: P = ρRT (1 + b ρ + b ρ2 + ...) • 1 2 1 2 van der Waals: P = RT (1/ρ b)− aρ • − − 2.4.7 Caloric equation of state 2.4.7.1 Description determined in experiments • gives e as function of ρ and T in general • arbitrary constant appears • must also be thermodynamically consistent via relation to be discussed later: •

CC BY-NC-ND. 28 October 2019, J. M. Powers. 2.5. SPECIAL CASES OF GOVERNING EQUATIONS 37

1 ∂P de = c (T ) dT T P dρ (2.118) v − ρ2 ∂T −   ρ !

With knowledge of cv(T ) and P (ρ, T ), the above can be integrated to find e.

2.4.7.2 Typical models

consistent with ideal gas: • – constant specific heat: e(T )= c (T T )+ e vo − o o – temperature dependent specific heat: e(T )= T c (Tˆ)dTˆ + e To v o

T R consistent with first virial: e(T )= cv(Tˆ)dTˆ + eo • To

R T consistent with van der Waals: e(ρ, T )= cv(Tˆ)dTˆ + a (ρ ρo)+ eo • To − − R 2.5 Special cases of governing equations

The governing equations are often expressed in more simple forms in common limits. Some are listed here.

2.5.1 One-dimensional equations Most of the mystery of vector notation is removed in the one-dimensional limit where v = w = 0, ∂ = ∂ = 0; additionally we adopt Stokes assumption λ = (2/3)µ: ∂y ∂z − ∂ρ ∂ρ ∂u + u + ρ = 0 (2.119) ∂t ∂x ∂x   ∂u ∂u ∂P ∂ 4 ∂u ρ + u = + µ + ρg (2.120) ∂t ∂x − ∂x ∂x 3 ∂x x     ∂e ∂e ∂ ∂T ∂u 4 ∂u 2 ρ + u = k P + µ (2.121) ∂t ∂x ∂x ∂x − ∂x 3 ∂x       µ = µ (ρ, T ) (2.122) k = k (ρ, T ) (2.123) P = P (ρ, T ) (2.124) e = e (ρ, T ) (2.125)

note: 7 equations, 7 unknowns: (ρ,u,P,e,T,µ,k)

CC BY-NC-ND. 28 October 2019, J. M. Powers. 38 CHAPTER 2. GOVERNING EQUATIONS

2.5.2 Euler equations When viscous stresses and heat conduction neglected, the Euler equations are obtained. dρ + ρ v = 0 (2.126) dt ∇· dv ρ = P (2.127) dt −∇ de d 1 = P (2.128) dt − dt ρ   e = e (P, ρ) (2.129)

Note:

6 equations, 6 unknowns (ρ,u,v,w,P,e) • body force neglected-usually unimportant in this limit • easy to show this is isentropic flow; energy change is all due to reversible P dv work •

Exercise: Write the one-dimensional Euler equations in a) non-conservative form, b) conservative form. Show all steps which lead from one form to the other.

2.5.3 Incompressible Navier-Stokes equations

If we take, ρ,k,µ,cp to be constant for an ideal gas and neglect viscous dissipation which is usually small in such cases:

v = 0 (2.130) ∇· dv ρ = P + µ 2v (2.131) dt −∇ ∇ dT ρc = k 2T (2.132) p dt ∇ Note:

5 equations, 5 unknowns: (u,v,w,P,T ) • mass and momentum uncoupled from energy • energy coupled to mass and momentum • detailed explanation required for use of c • p

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

Thermodynamics review

Suggested Reading:

Liepmann and Roshko, Chapter 1: pp. 1-24, 34-38 Shapiro, Chapter 2: pp. 23-44 Anderson, Chapter 1: pp. 12-25

As we have seen from the previous chapter, the subject of thermodynamics is a subset of the topic of viscous compressible flows. It is almost always necessary to consider the thermo- dynamics as part of a larger coupled system in design. This is in contrast to incompressible aerodynamics which can determine forces independent of the thermodynamics.

3.1 Preliminary mathematical concepts

If z = z(x, y) (3.1) then ∂z ∂z dz = dx + dy (3.2) ∂x ∂y y x which is of the form dz = Mdx + Ndy (3.3) Now ∂M ∂ ∂z = (3.4) ∂y ∂y ∂x ∂N ∂ ∂z = (3.5) ∂x ∂x ∂y ∂M ∂N thus = (3.6) ∂y ∂x so the implication is that if we are given dz,M,N, we can form z only if the above holds.

39 40 CHAPTER 3. THERMODYNAMICS REVIEW

3.2 Summary of thermodynamic concepts

property: characterizes the thermodynamics state of the system • – extensive: proportional to system’s mass, upper case variable E,S,H – intensive: independent of system’s mass, lower case variable e, s, h, (exceptions T,P )

equations of state: relate properties • Any intensive thermodynamic property can be expressed as a function of at most two • other intensive thermodynamic properties (for simple systems)

– P = ρRT : thermal equation of state for ideal gas

P – c = γ ρ : sound speed for calorically perfect ideal gas q first law: dEˆ = δQ δW • − second law: dS δQ/T • ≥ process: moving from one state to another, in general with accompanying heat transfer • and work

cycle: process which returns to initial state • reversible work: w = 2 P dv • 12 1 reversible heat transfer:R q = 2 Tds • 12 1 Figure 3.1 gives a sketch of an isothermalR thermodynamic process going from state 1 to state 2. The figure shows a variety of planes, P v, T s, P T , and v T . For ideal gases, 1) isotherms are hyperbolas in the P v plane:− P =(− RT )/v−, 2) isochores− are straight lines in the P T plane: P = (R/v)T , with− large v giving a small slope, and 3) isobars − are straight lines in the v T plane: v = (RT )/P , with large P giving a small slope. The area under the curve in the− P v plane gives the work. The area under the curve in the T s plane gives the heat transfer.− The energy change is given by the difference in the heat − transfer and the work. The isochores in the T s plane are non-trivial. For a calorically perfect ideal gas, they are given by exponential− curves. Figure 3.2 gives a sketch of a thermodynamic cycle. Here we only sketch the P v and − T s planes, though others could be included. Since it is a cyclic process, there is no net energy− change for the cycle and the cyclic work equals the cyclic heat transfer. The enclosed area in the P v plane, i.e. the net work, equals the enclosed area in the T s plane, i.e. the net heat transfer.− The sketch has the cycle working in the direction which− corresponds to an engine. A reversal of the direction would correspond to a refrigerator.

CC BY-NC-ND. 28 October 2019, J. M. Powers. 3.2. SUMMARY OF THERMODYNAMIC CONCEPTS 41

P T

v1

v2 P1

T1 = T 2 P 2 T1 = T 2

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

v2

P1 P 2 v1

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

Figure 3.1: Sketch of isothermal thermodynamic process

Example 3.1 Consider the following isobaric process for air, modelled as a calorically perfect ideal gas, from state 1 to state 2. P1 = 100 kPa, T1 = 300 K, T2 = 400 K.

Since the process is isobaric P = 100 kPa describes a straight line in P v and P T planes and P = P = 100 kPa. Since ideal gas, v T plane: − − 2 1 − R v = T straight lines! (3.7) P  

(287 J/kg/K)(300 K) v = RT /P = =0.861 m3/kg (3.8) 1 1 1 100, 000 Pa (287 J/kg/K)(400 K) v = RT /P = =1.148 m3/kg (3.9) 2 2 2 100, 000 Pa Since calorically perfect:

de = cvdT (3.10) e1 T1 de = cv dT (3.11) Ze2 ZT2 e e = c (T T ) (3.12) 2 − 1 v 2 − 1 = (716.5 J/kg/K)(400 K 300 K) (3.13) − = 71, 650 J/kg (3.14)

CC BY-NC-ND. 28 October 2019, J. M. Powers. 42 CHAPTER 3. THERMODYNAMICS REVIEW

P T

s v

q cycle = w cycle

Figure 3.2: Sketch of thermodynamic cycle

also

T ds = de + P dv (3.15)

T ds = cvdT + P dv (3.16) RT R RT from ideal gas : v = : dv = dT dP (3.17) P P − P 2 RT T ds = c dT + RdT dP (3.18) v − P dT dP ds = (c + R) R (3.19) v T − P dT dP ds = (c + c c ) R (3.20) v p − v T − P dT dP ds = c R (3.21) p T − P s2 T2 dT P2 dP ds = c R (3.22) p T − P Zs1 ZT1 ZP1 T P s s = c ln 2 R ln 2 (3.23) 2 − 1 p T − P  1   1  T P s s = c ln R ln (3.24) − o p T − P  o   o 

since P = constant: (3.25) T s s = c ln 2 (3.26) 2 − 1 p T  1  400 K = (1003.5 J/kg/K) ln (3.27) 300 K   = 288.7 J/kg/K (3.28)

v2 v2 w12 = P dv = P dv (3.29) Zv1 Zv1 = P (v v ) (3.30) 2 − 1

CC BY-NC-ND. 28 October 2019, J. M. Powers. 3.2. SUMMARY OF THERMODYNAMIC CONCEPTS 43

= (100, 000 Pa)(1.148 m3/kg 0.861 m3/kg) (3.31) − = 29, 600 J/kg (3.32)

Now

de = δq δw (3.33) − δq = de + δw (3.34) q = (e e )+ w (3.35) 12 2 − 1 12 q12 = 71, 650 J/kg + 29, 600 J/kg (3.36)

q12 = 101, 250 J/kg (3.37)

Now in this process the gas is heated from 300 K to 400 K. We would expect at a minimum that the surroundings were at 400 K. Let’s check for second law satisfaction.

q12 s2 s1 ? (3.38) − ≥ Tsurr 101, 250 J/kg 288.7 J/kg/K ? (3.39) ≥ 400 K 288.7 J/kg/K 253.1 J/kg/K yes (3.40) ≥

P T

v1

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

T1 = 300 K T1

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

P = P = 100 kPa v 1 2 2

v1

T T T T T 1 2 T 1 2

Figure 3.3: Sketch for isobaric example problem

CC BY-NC-ND. 28 October 2019, J. M. Powers. 44 CHAPTER 3. THERMODYNAMICS REVIEW

3.3 Maxwell relations and secondary properties

Recall 1 de = Tds Pd (3.41) − ρ   Since v 1/ρ we get ≡ de = Tds P dv (3.42) − Now we assume e = e(s, v), ∂e ∂e de = ds + dv (3.43) ∂s ∂v v s

Thus ∂e ∂e T = P = (3.44) ∂s − ∂v v s

and

∂T ∂2e ∂P ∂2e = = (3.45) ∂v ∂v∂s ∂s −∂s∂v s v

Thus we get a Maxwell relation : ∂T ∂P = (3.46) ∂v − ∂s s v

Define the following properties:

enthalpy: h e + pv • ≡ Helmholtz free energy: a e Ts • ≡ − Gibbs free energy: g h Ts • ≡ − Now with these definitions it is easy to form differential relations using the Gibbs relation as a root.

h = e + P v (3.47) dh = de + P dv + vdP (3.48) de = dh P dv vdP (3.49) − − substitute into Gibbs: de = Tds P dv (3.50) − dh P dv vdP = Tds P dv (3.51) − − − dh = Tds + vdP (3.52)

CC BY-NC-ND. 28 October 2019, J. M. Powers. 3.3. MAXWELL RELATIONS AND SECONDARY PROPERTIES 45

So s and P are natural variables for h. Through a very similar process we get the following relationships: ∂h ∂h = T = v (3.53) ∂s ∂P P s ∂a ∂a = P = s (3.54) ∂v − ∂T − T v ∂g ∂g = v = s (3.55) ∂P ∂T − T P

∂T ∂v ∂P ∂s ∂v ∂s = = = (3.56) ∂P ∂s ∂T ∂v ∂T − ∂P s P v T P T

The following thermodynamic properties are also useful and have formal definitions:

specific heat at constant volume: c ∂e • v ≡ ∂T v specific heat at constant pressure: c ∂h • p ≡ ∂T P ratio of specific heats: γ c /c • ≡ p v

∂P sound speed: c ∂ρ • ≡ s r 1 ∂v adiabatic compressibility : βs • ≡ − v ∂P s adiabatic bulk modulus: B v ∂P • s ≡ − ∂v s Generic problem: given P = P (T, v), find other properties

3.3.1 Internal energy from thermal equation of state Find the internal energy e(T, v) for a general material. e = e(T, v) (3.57) ∂e ∂e de = dT + dv (3.58) ∂T ∂v v T ∂e de = c dT + dv (3.59) v ∂v T

Now from Gibbs,

de = Tds P dv (3.60) − de ds = T P (3.61) dv dv − ∂e ∂s = T P (3.62) ∂v ∂v − T T

CC BY-NC-ND. 28 October 2019, J. M. Powers. 46 CHAPTER 3. THERMODYNAMICS REVIEW

Substitute from Maxwell relation, ∂e ∂P = T P (3.63) ∂v ∂T − T v

so ∂P de = c dT + T P dv (3.64) v ∂T −  v 

e T v ∂Pˆ deˆ = c (Tˆ)dTˆ + Tˆ Pˆ dvˆ (3.65) v ˆ eo To vo ∂T − ! Z Z Z vˆ

T v ∂Pˆ e(T, v)= e + c (Tˆ)dTˆ + Tˆ Pˆ dvˆ (3.66) o v ˆ To vo ∂T − ! Z Z vˆ

Example 3.2 Ideal gas

Find a general expression for e(T, v) if RT P (T, v)= (3.67) v Proceed as follows: ∂P = R/v (3.68) ∂T v

∂P RT T P = P (3.69) ∂T − v − v RT RT = =0 (3.70) v − v Thus e is T e(T )= eo + cv(Tˆ)dTˆ (3.71) ZTo We also find T h = e + P v = eo + cv(Tˆ)dTˆ + P v (3.72) ZTo T h(T, v)= eo + cv(Tˆ)dTˆ + RT (3.73) ZTo ∂h c (T, v)= = c (T )+ R = c (T ) (3.74) p ≡ ∂T v p P

R = cp(T ) cv(T ) (3.75) − Iff cv is a constant then e(T )= e + c (T T ) (3.76) o v − o h(T ) = (e + P v )+ c (T T ) (3.77) o o o p − o R = c c (3.78) p − v

CC BY-NC-ND. 28 October 2019, J. M. Powers. 3.3. MAXWELL RELATIONS AND SECONDARY PROPERTIES 47

Example 3.3 van der Waals gas

Find a general expression for e(T, v) if RT a P (T, v)= (3.79) v b − v2 − Proceed as before: ∂P R = (3.80) ∂T v b v − ∂P RT T P = P (3.81) ∂T − v b − v − RT RT a a = = (3.82) v b − v b − v2 v2 −  −  Thus e is T v a e(T, v)= e + c (Tˆ)dTˆ + dvˆ (3.83) o v vˆ2 ZTo Zvo T 1 1 = e + c (Tˆ)dTˆ + a (3.84) o v v − v ZTo  o  We also find T 1 1 h = e + P v = e + c (Tˆ)dTˆ + a + P v (3.85) o v v − v ZTo  o  T 1 1 RT v a h(T, v)= e + c (Tˆ)dTˆ + a + (3.86) o v v − v v b − v ZTo  o  − (3.87)

3.3.2 Sound speed from thermal equation of state Find the sound speed c(T, v) for a general material.

∂P c = (3.88) ∂ρ s s

2 ∂P c = (3.89) ∂ρ s

CC BY-NC-ND. 28 October 2019, J. M. Powers. 48 CHAPTER 3. THERMODYNAMICS REVIEW

Use Gibbs relation Tds = de + P dv (3.90) (3.91) Substitute earlier relation for de ∂P Tds = c dT + T P dv + P dv (3.92) v ∂T −   v   ∂P Tds = c dT + T dv (3.93) v ∂T v T ∂P Tds = c dT dρ (3.94) v − ρ2 ∂T ρ

Since P = P (T, v), P = P (T, ρ) ∂P ∂P dP = dT + dρ (3.95) ∂T ∂ρ ρ T

dP ∂P dρ ∂ρ dT = − T (3.96) ∂P ∂T ρ

Thus substituting for dT ∂P dP ∂ρ dρ T ∂P Tds = c − T dρ (3.97) v  ∂P  − ρ2 ∂T ∂T ρ ρ

  so grouping terms in dP and dρ we get

∂P c ∂ρ T ∂P Tds = v dP c T + dρ (3.98) ∂P −  v ∂P ρ2 ∂T  ∂T ρ ! ∂T ρ ρ

  SO if ds 0 we obtain ≡ ∂P ∂P 1 ∂P ∂ρ T ∂P = c T + (3.99) ∂ρ c ∂T  v ∂P ρ2 ∂T  s v ρ ∂T ρ ρ

 2 ∂P T ∂P = + (3.100) ∂ρ c ρ2 ∂T T v ρ!

So 2 ∂P T ∂P c(T, ρ)= + (3.101) v ∂ρ c ρ2 ∂T u T v ρ! u t CC BY-NC-ND. 28 October 2019, J. M. Powers. 3.3. MAXWELL RELATIONS AND SECONDARY PROPERTIES 49

Exercises: Liepmann and Roshko, 1.3 and 1.4, p. 383.

Example 3.4 Ideal gas

Find the sound speed if

P (T,ρ)= ρRT (3.102)

The necessary partials are

∂P ∂P = RT = ρR (3.103) ∂ρ ∂T T ρ

so

T 2 c(T,ρ)= RT + 2 (ρR) (3.104) s cvρ R2T = RT + (3.105) s cv R = RT 1+ (3.106) c s  v  c c = RT 1+ P − v (3.107) c s  v  c + c c = RT v P − v (3.108) c s  v  = γRT (3.109)

Sound speed depends on temperature alone for the caloricallyp perfect ideal gas.

Example 3.5 Virial gas

Find the sound speed if

P (T,ρ)= ρRT (1+ b1ρ) (3.110)

The necessary partials are

∂P ∂P = RT +2b ρRT = ρR (1 + b ρ) (3.111) ∂ρ 1 ∂T 1 T ρ

CC BY-NC-ND. 28 October 2019, J. M. Powers. 50 CHAPTER 3. THERMODYNAMICS REVIEW

so

T 2 c(T,ρ)= RT +2b1ρRT + 2 (ρR (1 + b1ρ)) (3.112) s cvρ R = RT 1+2b ρ + (1 + b ρ)2 (3.113) 1 c 1 s  v  Sound speed depends on both temperature and density.

Example 3.6 Thermodynamic process with a van der Waals Gas

A van der Waals gas with

R = 200 J/kg/K (3.114) a = 150 Pa m6/kg2 (3.115) b =0.001 m3/kg (3.116) c = [350 + 0.2(T 300K)] J/kg/K (3.117) v − 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 =1, 000 K. Find w13, q13, and s3 s1. Assume the surroundings are at 1, 000 K. Recall − RT a P = (3.118) v b − v2 − so at state 1 200 300 150 100, 000 = × (3.119) v 0.001 − v2 1 − 1 or expanding

0.15+ 150v 60, 100v2 + 100, 000v3 = 0 (3.120) − − Cubic equation–three solutions:

3 v1 =0.598 m /kg (3.121) 3 v1 =0.00125 0.0097im /kg not physical (3.122) − 3 v1 =0.00125+ 0.0097im /kg not physical (3.123)

Now at state 2 we know P2 and T2 so we can determine v2 200 300 150 1, 000, 000 = × (3.124) v 0.001 − v2 2 − 2 3 The physical solution is v2 =0.0585 m /kg. Now at state 3 we know v3 = v2 and T3. Determine P3: 200 1, 000 150 P = × =3, 478, 261 43, 831 = 3, 434, 430 Pa (3.125) 3 0.0585 0.001 − 0.05852 − − CC BY-NC-ND. 28 October 2019, J. M. Powers. 3.3. MAXWELL RELATIONS AND SECONDARY PROPERTIES 51

Now w = w + w = 2 P dv + 3 P dv = 2 P dv since 2 3 is at constant volume. So 13 12 23 1 2 1 − R R R v2 RT a w = dv (3.126) 13 v b − v2 Zv1  −  v2 dv v2 dv = RT a (3.127) 1 v b − v2 Zv1 − Zv1 v b 1 1 = RT ln 2 − + a (3.128) 1 v b v − v  1 −   2 1  0.0585 0.001 1 1 = 200 300 ln − + 150 (3.129) × 0.598 0.001 0.0585 − 0.598  −    = 140, 408+2, 313 (3.130) − = 138, 095 J/kg = 138 kJ/kg (3.131) − − The gas is compressed, so the work is negative. Since e is a state property:

T3 1 1 e e = c (T )dT + a (3.132) 3 − 1 v v − v ZT1  1 3  Now 1 c = 350+0.2(T 300) = 290+ T (3.133) v − 5 so

T3 1 1 1 e e = 290 + T dT + a (3.134) 3 − 1 5 v − v ZT1    1 3  1 1 1 = 290(T T )+ T 2 T 2 + a (3.135) 3 − 1 10 3 − 1 v − v  1 3  1  1 1 290 (1, 000 300) + 1, 0002 3002 + 150 (3.136) − 10 − 0.598 − 0.0585   = 203, 000+ 91, 000 2, 313 (3.137) − = 291, 687 J/kg = 292 kJ/kg (3.138)

Now from the first law

e e = q w (3.139) 3 − 1 13 − 13 q = e e + w (3.140) 13 3 − 1 13 q = 292 138 (3.141) 13 − q13 = 154 kJ/kg (3.142)

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

Now find the entropy change. Manipulate the Gibbs equation:

T ds = de + P dv (3.143) 1 P ds = de + dv (3.144) T T 1 a P ds = c (T )dT + dv + dv (3.145) T v v2 T   CC BY-NC-ND. 28 October 2019, J. M. Powers. 52 CHAPTER 3. THERMODYNAMICS REVIEW

1 a 1 RT a ds = c (T )dT + dv + dv (3.146) T v v2 T v b − v2  −    c (T ) R ds = v dT + dv (3.147) T v b − T3 c (T ) v b s s = v dT + R ln 3 − (3.148) 3 − 1 T v b ZT1 1 − 1,000 290 1 v b = + dT + R ln 3 − (3.149) T 5 v b Z300   1 − 1, 000 1 0.0585 0.001 = 290 ln + (1, 000 300) + 200 ln − (3.150) 300 5 − 0.598 0.001 − = 349+140 468 (3.151) − J kJ = 21 =0.021 (3.152) kg K kg K Is the second law satisfied for each portion of the process?

First look at 1 2 → e e = q w (3.153) 2 − 1 12 − 12 q = e e + w (3.154) 12 2 − 1 12 T2 1 1 v b 1 1 q = c (T )dT + a + RT ln 2 − + a (3.155) 12 v v − v 1 v b v − v ZT1  1 2 !   1 −   2 1  (3.156)

Since T1 = T2 and canceling the terms in a we get

v b 0.0585 0.001 J q = RT ln 2 − = 200 300 ln − = 140, 408 (3.157) 12 1 v b × 0.598 0.001 − kg  1 −   −  Since isothermal v b s s = R ln 2 − (3.158) 2 − 1 v b  1 −  0.0585 0.001 = 200 ln − (3.159) 0.598 0.001  −  J = 468.0 (3.160) − 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 12 ? (3.161) 2 − 1 ≥ T J J 140, 408 468.0 − kg (3.162) − kg K ≥ 300 K J J 468.0 468.0 ok (3.163) − kg K ≥− kg K

CC BY-NC-ND. 28 October 2019, J. M. Powers. 3.4. CANONICAL EQUATIONS OF STATE 53

Next look at 2 3 → q = e e + w (3.164) 23 3 − 2 23 T3 1 1 v3 q = c (T )dT + a + P dv (3.165) 23 v v − v ZT2  2 3 ! Zv2  T3 since isochoric q23 = cv(T )dT (3.166) ZT2 1000 T J = 290 + dT = 294, 000 (3.167) 5 K Z300   Now look at the entropy change for the isochoric process:

T3 c (T ) s s = v dT (3.168) 3 − 2 T ZT2 T3 290 1 = + dT (3.169) T 5 ZT2   1, 000 1 J = 290 ln + (1, 000 300) = 489 (3.170) 300 5 − 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 1, 000 K to assess the influence≥ of the second law. q s s 23 ? (3.171) 3 − 2 ≥ T J J 294, 000 489 kg (3.172) kg K ≥ 1, 000 K J J 489 294 ok (3.173) kg K ≥ kg K

3.4 Canonical equations of state

If we have a single equation of state in a special canonical form, we can form both thermal and caloric equations. Since de = Tds P dv (3.174) − dh = Tds + vdP (3.175) it is suggested that the form e = e(s, v) (3.176) h = h(s,P ) (3.177)

CC BY-NC-ND. 28 October 2019, J. M. Powers. 54 CHAPTER 3. THERMODYNAMICS REVIEW

is useful.

Example 3.7 Canonical Form

If s h(s, P )= Kc P R/cp exp + (h c T ) (3.178) p c o − p o  p  derive both thermal and caloric state equations P (v,T ) and e(v,T ).

Now for our material ∂h s = KP R/cp exp (3.179) ∂s c P  p  ∂h − s = KRP R/cp 1 exp (3.180) ∂P c s  p 

Now since ∂h = T (3.181) ∂s P ∂h = v (3.182) ∂P s

we have s T = KP R/cp exp (3.183) c  p  − s v = KRP R/cp 1 exp (3.184) c  p  Dividing one by the other gives T P = (3.185) v R RT P = (3.186) v Substituting our expression for T into our canonical equation for h we also get

h = c T + (h c T ) (3.187) p o − p o h = c (T T )+ h (3.188) p − o o

which is useful in itself. Substituting in for T and To

P v P v h = c o o + h (3.189) p R − R o   Using h e + P v we get ≡ P v P v e + P v = c o o + e + P v (3.190) p R − R o o o   CC BY-NC-ND. 28 October 2019, J. M. Powers. 3.5. ISENTROPIC RELATIONS 55

so c c e = p 1 P v p 1 P v + e (3.191) R − − R − o o o  c   e = p 1 (P v P v )+ e (3.192) R − − o o o  c  e = p 1 (RT RT )+ e (3.193) R − − o o e = (c  R) (T T )+ e (3.194) p − − o o e = [c (c c )] (T T )+ e (3.195) p − p − v − o o e = c (T T )+ e (3.196) v − o o So one canonical equation gives us all the information we need! Oftentimes, it is difficult to do a single experiment to get the canonical form.

Exercise: For a calorically perfect ideal gas, write the Helmholtz free energy and Gibbs free energy in canonical form, i.e. what is a(T, v), g(P, T )?

3.5 Isentropic relations

Of particular importance in thermodynamics in general and compressible flow in particular are relations that describe an , s = constant. Recall the second law. δq ds (3.197) ≥ T If the process is reversible, δq ds = (3.198) T If the process is adiabatic δq 0 so ds = 0 (3.199) ≡ So an isentropic process is both adiabatic and reversible. We know from the first law written in terms of entropy that this implies that q 0 (3.200) ≡ τ 0 (3.201) ≡ In this case the Gibbs relation and the first law reduce to the same expression: de = P dv (3.202) − That is the energy change is all due to reversible pressure volume work.

We would like to develop an expression between two variables for an isentropic process.

With knowledge of P (T, v)

CC BY-NC-ND. 28 October 2019, J. M. Powers. 56 CHAPTER 3. THERMODYNAMICS REVIEW

form e(T, v) • eliminate T to form e(P, v) • take derivative and substitute into Gibbs/First Law •

∂e ∂e dP + dv = P dv (3.203) ∂P ∂v − v P ∂e ∂e dP + + P dv = 0 (3.204) ∂P ∂v v  P 

Integration of this equation gives a relationship between P and v.

Example 3.8 Calorically Perfect Ideal Gas

Find the relationship for a calorically perfect ideal gas which undergoes an isentropic process.

Ideal Gas: P v = RT (3.205) Calorically Perfect: e = cvT + eo (3.206) Thus P v c 1 e = c + e = v P v + e = P v + e (3.207) v R o c c o γ 1 o P − v − Thus the necessary derivatives are ∂e 1 = v (3.208) ∂P γ 1 v − ∂e 1 = P (3.209) ∂v γ 1 P −

so substituting into our developed relationship gives 1 1 vdP + P + P dv = 0 (3.210) γ 1 γ 1 −  −  vdP + γP dv = 0 (3.211) dP dv = γ (3.212) P − v P v ln = γ ln (3.213) Po − vo P v γ ln = ln o (3.214) Po v P v γ = o (3.215) Po v γ γ   P v = Povo = constant (3.216)

CC BY-NC-ND. 28 October 2019, J. M. Powers. 3.5. ISENTROPIC RELATIONS 57

also using the thermal state equation

RT γ P v T vo vo = RT = = (3.217) Po o To v v vo  γ−1 T v γ−1 P γ = o = (3.218) To v Po     Find the work in a process from v1 to v2

1 w12 = P dv (3.219) Z2 v2 dv = P vγ (3.220) o o vγ Zv1 v1−γ v2 = P vγ (3.221) o o 1 γ  v1 γ − P v − − = o o v1 γ v1 γ (3.222) 1 γ 2 − 1 −  P v P v = 2 2 − 1 1 (3.223) 1 γ − Also

de = δq δw =0 δw so (3.224) − − P v P v e e = 2 2 − 1 1 (3.225) 2 − 1 γ 1 −

Figure 3.4 gives a sketch for the calorically perfect ideal gas undergoing an isentropic expansion in various planes.

Example 3.9 Virial Gas

Find the relationship between P and v for a virial gas with constant cv which undergoes an isentropic process.

Virial Gas: RT P = (3.226) v b − This is van der Waals with a = 0 and cv constant so:

e = cvT + eo (3.227)

Thus P (v b) e = c − + e (3.228) v R o

CC BY-NC-ND. 28 October 2019, J. M. Powers. 58 CHAPTER 3. THERMODYNAMICS REVIEW

P T

v1

P1 v2

T1 T 1 T2

P2 T2 v v s = s s 1 2 v 1 2 2 2 w = ∫ P dv e - e = q - w q = ∫ T ds 12 1 2 1 12 12 12 1 P v P2

v1

v2 P1

v P1 v2 1

P2 T T T T T 2 1 T 2 1

Figure 3.4: Sketch of isentropic expansion process

Thus the necessary derivatives are

∂e c = v (v b) (3.229) ∂P R − v ∂e c = v P (3.230) ∂v R P

so substituting into our developed relationship gives c c v (v b) dP + v P + P dv = 0 (3.231) R − R  R  (v b) dP + 1+ P dv = 0 (3.232) − c  v  R withγ ˆ 1+ (3.233) ≡ cv dP γˆ + P = 0 (3.234) dv v b − γˆ dP γˆ γˆ exp dv + exp dv P = 0 (3.235) v b dv v b v b  Z −   Z −  − dP γˆ exp ln (v b)γˆ + exp ln (v b)γˆ P = 0 (3.236) − dv − v b     dP  −γˆ (v b)γˆ + (v b)γˆ P = 0 (3.237) − dv − v b d − (v b)γˆ P = 0 (3.238) dv −   CC BY-NC-ND. 28 October 2019, J. M. Powers. 3.5. ISENTROPIC RELATIONS 59

(v b)γˆ P = (v b)γˆ P (3.239) − o − o P v b γˆ = o − (3.240) P v b o  − 

Exercise: Find the relationship between T and v for a virial gas in an isentropic process.

Exercise: Find an expression for the work done by a van der Waals gas in an isentropic process.

J m3 Exercise: A virial gas, m =3 kg with R = 290 kgK , b =0.002 kg with constant specific kJ heat cv = 0.700 kg K is initially at P = 1.2 bar and T = 320 K. It undergoes a two step process: 1 2 is an isochoric compression to 500 kP a; 2 3 is an isentropic expansion to → → 300 kP a. Find the total work W13 in units of J, the total heat transfer Q13 in units of J, and the change in entropy S S in units of J/K. Include a sketch, roughly to scale, of the 3 − 1 total process in the P v and T s planes. − −

CC BY-NC-ND. 28 October 2019, J. M. Powers. 60 CHAPTER 3. THERMODYNAMICS REVIEW

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

One-dimensional compressible flow

White, Chapter 9: pp. 511-559, Liepmann and Roshko, Chapter 2: pp. 39-65, Hughes and Brighton, Chapter 7: pp. 178-185, Shapiro, Vol. 1, Chapters 4-8: pp. 73-262,

This chapter will discuss one-dimensional flow of a compressible fluid. Notation can pose problems, and many common ones are in use. Here a new convention will be adopted. In this chapter

velocity in the x-direction will be denoted as u, • specific internal energy, denoted in previous chapters by u, will here be e, • total internal energy, denoted in previous chapters by U, will here be E. • The following topics will be covered:

development of generalized one-dimensional flow equations, • isentropic flow with area change, • flow with normal shock waves, • flow with friction (Fanno flow), • flow with heat transfer (Rayleigh flow), • flow in a shock tube. •

Assume for this chapter:

The flow is uni-directional in the x direction with u = 0 and with the y and z • components of the velocity vector both− zero: v 0,w 6 0 − − ≡ ≡ 61 62 CHAPTER 4. ONE-DIMENSIONAL

Spatial gradients are admitted in x, but not in y or z: ∂ = 0, ∂ 0, ∂ 0. • ∂x 6 ∂y ≡ ∂z ≡ Friction and heat transfer will not be modelled rigorously. Instead, they will be modelled in a fashion which captures the relevant physics and retains analytic tractability.

4.1 Generalized one-dimensional equations

Flow with area change is illustrated by the following sketch of a control volume:. See Figure 4.1.

ρ 1 ρ 2 u1 u2 A1 A2 P1 P2 e1 τ e2 x1 w

q x2 x - x = ∆ x Perimeter = L 2 1

Figure 4.1: Control volume sketch

For this problem adopt the following conventions

surface 1 and 2 are open and allow fluxes of mass, momentum, and energy • surface w is a closed wall; no mass flux through the wall • W external heat flux qw (Energy/Area/Time: m2 ) through the wall allowed-qw known fixed • parameter

diffusive, longitudinal heat transfer ignored, q =0 • x N wall shear τ (Force/Area: 2 ) allowed–τ known, fixed parameter • w m w diffusive viscous stress not allowed τ =0 • xx cross-sectional area a known fixed function: A(x) •

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.1. GENERALIZED ONE-DIMENSIONAL EQUATIONS 63

4.1.1 Mass Take the overbar notation to indicate a volume averaged quantity.

The amount of mass in a control volume after a time increment ∆t is equal to the original amount of mass plus that which came in minus that which left:

ρ¯A¯∆x =ρ ¯A¯∆x + ρ A (u ∆t) ρ A (u ∆t) (4.1) t+∆t t 1 1 1 − 2 2 2

Rearrange and divide by ∆x∆t:

ρ¯A¯ ρ¯A¯ ρ A u ρ A u t+∆t − t + 2 2 2 − 1 1 1 = 0 (4.2) ∆t ∆x (4.3)

Taking the limit as ∆t 0, ∆x 0: → → ∂ ∂ (ρA)+ (ρAu) = 0 (4.4) ∂t ∂x If steady

d (ρAu) = 0 (4.5) dx dρ dA du Au + ρu + ρA = 0 (4.6) dx dx dx 1 dρ 1 dA 1 du + + = 0 (4.7) ρ dx A dx u dx

Integrate from x1 to x2: x2 d x2 (ρAu)dx = 0dx (4.8) x1 dx x1 Z 2 Z d (ρAu) = 0 (4.9) Z1 ρ u A ρ u A = 0 (4.10) 2 2 2 − 1 1 1 ρ u A = ρ u A m˙ = C (4.11) 2 2 2 1 1 1 ≡ 1 4.1.2 Momentum Newton’s Second Law says the time rate of change of linear momentum of a body equals the sum of the forces acting on the body. In the x direction this is roughly as follows:

d (mu)= F (4.12) dt x

CC BY-NC-ND. 28 October 2019, J. M. Powers. 64 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

In discrete form this would be mu mu |t+∆t − |t = F (4.13) ∆t x mu = mu + F ∆t (4.14) |t+∆t |t x For a control volume containing fluid, one must also account for the momentum which enters and leaves the control volume. The amount of momentum in a control volume after a time increment ∆t is equal to the original amount of momentum plus that which came in minus that which left plus that introduced by the forces acting on the control volume.

pressure force at surface 1 pushes fluid • pressure force at surface 2 restrains fluid • force due to the reaction of the wall to the pressure force pushes fluid if area change • positive

force due to the reaction of the wall to the shear force restrains fluid •

¯ ¯ ρ¯A∆x u¯ t+∆t = ρ¯A∆x u¯ t  +(ρ1A1 (u1∆t)) u1

(ρ A (u ∆t)) u − 2 2 2 2 +(P A ) ∆t (P A ) ∆t 1 1 − 2 2 + P¯ (A A ) ∆t 2 − 1 τ ¯∆x ∆t − wL  Rearrange and divide by ∆x∆t: 

ρ¯A¯u¯ ρ¯A¯u¯ ρ A u2 ρ A u2 t+∆t − t + 2 2 2 − 1 1 1 ∆t ∆x P A P A A A = 2 2 − 1 1 + P¯ 2 − 1 τ ¯ − ∆x ∆x − wL In the limit ∆x 0, ∆t 0 one gets → → ∂ ∂ ∂ ∂A (ρAu)+ ρAu2 = (P A)+ P τ (4.15) ∂t ∂x −∂x ∂x − wL  In steady state: d d dA ρAu2 = (P A)+ P τ (4.16) dx −dx dx − wL  CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.1. GENERALIZED ONE-DIMENSIONAL EQUATIONS 65

du d dA dP dA ρAu + u (ρAu) = P A + P τ (4.17) dx dx − dx − dx dx − wL du dP ρu = τ L (4.18) dx − dx − w A ρudu + dP = τ Ldx (4.19) − w A 1 du + dP = τ L dx (4.20) ρu − w m˙ u2 ρd + dP = τ Ldx (4.21) 2 − w A   Wall shear lowers the combination of pressure and dynamic head. If no wall shear: u2 dP = ρd (4.22) − 2   Increase in velocity magnitude decreases the pressure. If no area change dA = 0 and no friction τ 0: w ≡ du dP ρu + = 0 (4.23) dx dx d add u mass u (ρu) = 0 (4.24) dx d ρu2 + P = 0 (4.25) dx 2 2 ρu + P = ρouo + Po = C2 (4.26)

4.1.3 Energy The first law of thermodynamics states that the change of total energy of a body equals the heat transferred to the body minus the work done by the body:

E E = Q W (4.27) 2 − 1 − E = E + Q W (4.28) 2 1 − So for the control volume this becomes the following when one also accounts for the energy flux in and out of the control volume in addition to the work and heat transfer: u¯2 u¯2 ρ¯A¯∆x e¯ + = ρ¯A¯∆x e¯ + 2 2   t+∆t   t   2 2 u1 u2 +ρ1A1 (u1∆t) e1 + ρ2A2 (u2∆t) e2 + 2 − 2     +q ¯∆x ∆t +(P A )(u ∆t) (P A )(u ∆t) w L 1 1 1 − 2 2 2  CC BY-NC-ND. 28 October 2019, J. M. Powers. 66 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

Note: mean pressure times area difference does no work because acting on stationary bound- • ary work done by shear force not included1 • Rearrange and divide by ∆t∆x:

¯ u¯2 ¯ u¯2 ρ¯A e¯ + 2 ρ¯A e¯ + 2 t+∆t − t     ∆t u2 u2 ρ A u e + 2 + P2 ρ A u e + 1 + P1 2 2 2 2 2 ρ2 1 1 1 1 2 ρ1 + − = q ¯  ∆x   wL In differential form as ∆x 0, ∆t 0 → → ∂ u2 ∂ u2 P ρA e + + ρAu e + + = q ∂t 2 ∂x 2 ρ wL       In steady state: d u2 P ρAu e + + = q (4.29) dx 2 ρ wL    d u2 P u2 P d ρAu e + + + e + + (ρAu) = q (4.30) dx 2 ρ 2 ρ dx wL     d u2 P q ρu e + + = wL (4.31) dx 2 ρ A   de du 1 dP P dρ q ρu + u + = wL (4.32) dx dx ρ dx − ρ2 dx A   subtract product of momentum and velocity (4.33) du dP τ u ρu2 + u = wL (4.34) dx dx − A de Pu dρ q τ u ρu = wL + wL (4.35) dx − ρ dx A A de P dρ (q + τ u) = w w L (4.36) dx − ρ2 dx m˙ 1In neglecting work done by the wall shear force, I have taken an approach which is nearly universal, but fundamentally difficult to defend. At this stage of the development of these notes, I am not ready to enter into a grand battle with all established authors and probably confuse the student; consequently, results for flow with friction will be consistent with those of other sources. The argument typically used to justify this is that the real fluid satisfies no-slip at the boundary; thus, the wall shear actually does no work. However, one can easily argue that within the context of the one-dimensional model which has been posed that the shear force behaves as an external force which reduces the fluid’s mechanical energy. Moreover, it is possible to show that neglect of this term results in the loss of frame invariance, a serious defect indeed. To model the work of the wall shear, one would include the term τ ¯∆x (¯u∆t) in the energy equation. w L  CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.1. GENERALIZED ONE-DIMENSIONAL EQUATIONS 67

Since e = e(P, ρ)

∂e ∂e de = dρ + dP (4.37) ∂ρ ∂P P ρ

de ∂e dρ ∂e dP = + (4.38) dx ∂ρ dx ∂P dx P ρ

so ∂e dρ ∂e dP P dρ (q + τ u) + = w w L (4.39) ∂ρ dx ∂P dx − ρ2 dx m˙ P ρ

∂e P dP ∂ρ ρ2 dρ (q + τ u) + P − = w w L (4.40) dx  ∂e  dx ∂e ∂P ρ m˙ ∂P ρ

  Now it can be shown that

∂e P ∂P ∂ρ ρ2 c2 = = P − (4.41) ∂ρ −  ∂e  s ∂P ρ

  so dP dρ (q + τ u) c2 = w w L (4.42) dx − dx ∂e m˙ ∂P ρ

dP 2 dρ (qw + τw u) c = L (4.43) dx − dx ∂e ρuA ∂P ρ

Special case of flow with no heat transfer qw 0. Area change allowed!, wall friction allowed! (see earlier footnote): ≡

d u2 P ρu e + + = 0 (4.44) dx 2 ρ  2  2 u P uo Po e + + = eo + + = C3 (4.45) 2 ρ 2 ρo u2 u2 h + = h + o = C (4.46) 2 o 2 3

Example 4.1 Adiabatic Flow of Argon2

2adopted from White’s 9.1, p. 583

CC BY-NC-ND. 28 October 2019, J. M. Powers. 68 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

5 Given: Argon, γ = 3 , flows adiabatically through a duct. At section 1, P1 = 200 psia, T1 = ◦ ft ft 500 F,u1 = 250 s . At section 2 P2 = 40 psia, u2 =1, 100 s .

◦ Btu Find: T in F and s s in ◦ 2 2 − 1 lbm R ft lbf Btu Assume: Ar is a calorically perfect ideal gas, tables give R = 38.68 lbmR ,cp =0.1253 lbmR

Analysis: First get the units into shape:

T1 =500+460 = 960 R (4.47) Btu ft lbf lbm ft ft2 c = 0.1253 779 32.17 = 3, 140 (4.48) p lbm R Btu lbf s2 s2 R       ft lbf lbm ft ft2 R = 38.68 32.17 = 1, 244 (4.49) lbm R lbf s2 s2 R     ft lbf 1 Btu Btu R = 38.68 = 0.04965 (4.50) lbm R 779 ft lbf lbm R     Now consider an energy balance: u2 u2 h + 2 = h + 1 (4.51) 2 2 1 2 u2 u2 c T + h + 2 = c T + h + 1 (4.52) p 2 o 2 p 1 o 2 1 2 2 T2 = T1 + u1 u2 (4.53) 2cp − 2 2  1 1 ft ft T = 960 R + 250 1, 100 = 777 R (4.54) 2 2 ft2 s − s 3, 140 s2R     ! ◦ T = 777 460 = 317 F (4.55) 2 − The flow sped up; temperature went down. was converted into kinetic energy

Calculate the entropy change. For the calorically perfect ideal gas: T P s s = c ln 2 R ln 2 (4.56) 2 − 1 p T − P  1   1  Btu 777 R Btu 40 psia = 0.1253 ln 0.04965 ln (4.57) lbm R 960 R − lbm R 200 psia     Btu = 0.0265 ( .0799) = 0.0534 (4.58) − − − lbm R Entropy change positive. Since adiabatic, there must have been irreversible friction which gave rise to this.

Example 4.2 Adiabatic Flow of Steam3 3adopted from White’s 9.2, p. 583

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.1. GENERALIZED ONE-DIMENSIONAL EQUATIONS 69

Same problem now with steam Given: Steam flows adiabatically through a duct. At section 1, ◦ ft ft P1 = 200 psia, T1 = 500 F,u1 = 250 s . At section 2 P2 = 40 psia, u2 =1, 100 s .

◦ Btu Find: T in F and s s in ◦ 2 2 − 1 lbm R Analysis: Use steam tables for property values. Energy balance:

u2 u2 h + 2 = h + 1 (4.59) 2 2 1 2 1 h = h + u2 u2 (4.60) 2 1 2 1 − 2 Btu 1  1 Btu 1 lbf s2 ft 2 ft 2 h = 1269 + 250 1, 100 (4.61) 2 lbm 2 779 ft lbf 32.17 lbm ft s − s         ! Btu h = 1, 246 (4.62) 2 lbm

Btu Interpolate steam tables at P2 = 40 psia, h2 = h2 =1, 246 lbm and find ◦ T2 = 420 F (4.63) Btu s = 1.7720 (4.64) 2 lbm R

Btu Tables give s1 =1.6239 lbm R so the entropy change is Btu s s =1.7720 1.6239 = 0.148 (4.65) 2 − 1 − lbm R

Example 4.3 Flow of Air with Heat Addition

m Given: Air initially at P1 = 100 kPa, T1 = 300 K, u1 = 10 s flows in a duct of length 100 m. The duct has a constant circular cross sectional area of A = 0.02 m2 and is isobarically heated with a constant heat flux qw along the entire surface of the duct. At the end of the duct the flow has P2 = 100 kPa, T2 = 500 K

Find: the mass flow ratem ˙ , the wall heat flux qw and the entropy change s2 s1; check for satisfaction of the second law. −

kJ kJ Assume: Calorically perfect ideal gas, R =0.287 kg K , cp =1.0035 kg K

Analysis:

CC BY-NC-ND. 28 October 2019, J. M. Powers. 70 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

Geometry:

A = πr2 (4.66) A r = (4.67) r π =2πr =2√πA = 2 π (0.02 m2)=0.501 m (4.68) L p Get the mass flux.

P1 = ρ1RT1 (4.69) P1 100 kPa ρ1 = = (4.70) RT1 kJ 0.287 kg K (300 K)  kg  = 1.161 (4.71) m3 So kg m kg m˙ = ρ u A = 1.161 10 0.02 m2 =0.2322 (4.72) 1 1 1 m3 s s      Get the flow variables at state 2:

P2 100 kPa ρ2 = = (4.73) RT2 kJ 0.287 kg K (500 K)  kg  = 0.6969 (4.74) m3 ρ2u2A2 = ρ1u1A1 (4.75) ρ1u1A1 ρ1u1 u2 = = (4.76) ρ2A2 ρ2 kg m 1.161 m3 10 s m = = 16.67 (4.77)   kg s 0.6969 m 3  Now consider the energy equation:

d u2 P q ρu e + + = wL (4.78) dx 2 ρ A   d u2 q h + = wL (4.79) dx 2 m˙   L d u2 L q h + dx = wLdx (4.80) dx 2 m˙ Z0   Z0 u2 u2 q L h + 2 h 1 = w L (4.81) 2 2 − 1 − 2 m˙ u2 u2 q L c (T T )+ 2 1 = w L (4.82) p 2 − 1 2 − 2 m˙ m˙ u2 u2 q = c (T T )+ 2 1 (4.83) w L p 2 − 1 2 − 2  L   (4.84)

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.1. GENERALIZED ONE-DIMENSIONAL EQUATIONS 71

Substituting the numbers, one finds,

kg m 2 m 2 0.2322 s J 16.67 s 10 s qw = 1, 003.5 (500 K 300 K)+ (4.85) (100 m) (0.501 m)! kg K − 2  − 2  ! kg J m2 q = 0.004635 200, 700 88.9 (4.86) w m2 s kg − s2   kg J J q = 0.004635 200, 700 88.9 (4.87) w m2 s kg − kg   W q = 930 (4.88) w m2 Heat flux positive, denoting heat into the air.

Now find the entropy change.

T P s s = c ln 2 R ln 2 (4.89) 2 − 1 p T − P  1   1  J 500 K J 100 kPa s s = 1, 003.5 ln 287 ln (4.90) 2 − 1 kg K 300 K − kg K 100 kPa         J s s = 512.6 0 = 512.6 (4.91) 2 − 1 − kg K

Is the second law satisfied? Assume the heat transfer takes place from a reservoir held at 500 K. The reservoir would have to be at least at 500 K in order to bring the fluid to its final state of 500 K. It could be greater than 500 K and still satisfy the second law.

Q S S 12 (4.92) 2 − 1 ≥ T Q˙ S˙ S˙ 12 (4.93) 2 − 1 ≥ T Q˙ m˙ (s s ) 12 (4.94) 2 − 1 ≥ T q A m˙ (s s ) w tot (4.95) 2 − 1 ≥ T q L m˙ (s s ) w L (4.96) 2 − 1 ≥ T q L s s w L (4.97) 2 − 1 ≥ mT˙ J J 930 2 (100 m) (0.501 m) 512.6 s m (4.98) kg K ≥ kg 0.2322 s (500 K) J  J  512.6 401.3 (4.99) kg K ≥ kg K

CC BY-NC-ND. 28 October 2019, J. M. Powers. 72 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

4.1.4 Influence coefficients Now, uncouple these equations. First, summarize: dρ du ρu dA u + ρ = (4.100) dx dx − A dx du dP τ ρu + = wL (4.101) dx dx − A dP dρ (q + τ u) c2 = w w L (4.102) dx − dx ∂e ρuA ∂P ρ

In matrix form this is ρu dA dρ u ρ 0 A dx dx − τ du w 0 ρu 1 = AL (4.103) dx  (qw−+τwu)   2   dP  L c 0 1 dx ρuA ∂e −  ∂P |ρ        Use Cramer’s Rule to solve for the derivatives. First calculate the determinant of the coef- ficient matrix: u ((ρu)(1) (1)(0)) ρ (0)(1) ( c2)(1) = ρ u2 c2 (4.104) − − − − − Implementing Cramer’s Rule:  

ρu dA τw (qw+τw u) ρu ρ L + ρ L A dx A ρuA ∂e dρ − − − ∂P =  |ρ  (4.105) dx  ρ(u2 c2) − 2 ρu dA τw (qw+τwu) c + u L u L A dx A ρuA ∂e du − − − − ∂P ρ =  |  (4.106) dx  ρ (u2 c2) − 2 ρu dA 2 τw 2 (qw+τwu) ρuc ρc L + ρu L A dx A ρuA ∂e dP − − − ∂P =  |ρ  (4.107) dx  ρ (u2 c2) − Simplify

2 dA (qw+τwu) ρu + τ + L dx w ρu ∂e dρ 1 − L ∂P = |ρ (4.108) dx A (u2 c2) − 2 dA (qw+τwu) c ρu uτ L dx w ρ ∂e du 1 − L − ∂P = |ρ (4.109) dx A ρ (u2 c2) − 2 2 dA 2 (qw+τwu) u c ρu + c τ + L dx w ρ ∂e dP 1 − L ∂P = |ρ (4.110) dx A (u2 c2) − Note:

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.2. FLOW WITH AREA CHANGE 73

a system of coupled non-linear ordinary differential equations • in standard form for dynamic system analysis: du = f(u) • dx valid for general equations of state • singular when velocity sonic u = c • 4.2 Flow with area change

This section will consider flow with area change with an emphasis on isentropic flow. Some problems will involve non-isentropic flow but a detailed discussion of such flows will be delayed.

4.2.1 Isentropic Mach number relations Take special case of

τ =0 • w q =0 • w calorically perfect ideal gas (CPIG) • Then d (ρuA) = 0 (4.111) dx du dP ρu + = 0 (4.112) dx dx d u2 P e + + = 0 (4.113) dx 2 ρ   Integrate the energy equation with h = e + P/ρ

u2 u2 h + = h + o (4.114) 2 o 2

If one defines the “o” condition to be a condition of rest, then uo 0. This is a stagnation condition. So ≡ u2 h + = h (4.115) 2 o u2 (h h )+ = 0 (4.116) − o 2

CC BY-NC-ND. 28 October 2019, J. M. Powers. 74 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

Since CPIG,

u2 c (T T )+ = 0 (4.117) p − o 2 u2 T To + = 0 (4.118) − 2cp T u2 1 o + = 0 (4.119) − T 2cpT Now note that

cp cv cp cp cv γR cp = cp − = cp − = (4.120) cp cv cv 1 γ 1 − cv − − so T γ 1 u2 1 o + − = 0 (4.121) − T 2 γRT T γ 1 u2 o = 1+ − (4.122) T 2 γRT Recall the sound speed and Mach number for a CPIG:

2 c = γRT if P = ρRT, e = cvT + eo (4.123) u 2 M 2 (4.124) ≡ c   thus, T γ 1 o = 1+ − M 2 (4.125) T 2 1 T γ 1 − = 1+ − M 2 (4.126) T 2 o   Now if the flow is isentropic one has

− γ 1 γ 1 T ρ − P γ = = (4.127) T ρ P o  o   o  Thus

1 ρ γ 1 − γ−1 = 1+ − M 2 (4.128) ρo 2   γ P γ 1 − γ−1 = 1+ − M 2 (4.129) P 2 o   CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.2. FLOW WITH AREA CHANGE 75

For air γ =7/5 so

1 T 1 − = 1+ M 2 (4.130) To 5   5 ρ 1 − 2 = 1+ M 2 (4.131) ρo 5   7 P 1 − 2 = 1+ M 2 (4.132) P 5 o   Figures 4.2, 4.3 4.4 show the variation of T , ρ and P with M 2 for isentropic flow.

Other thermodynamic properties can be determined from these, e.g. sound speed:

1/2 c γRT T γ 1 − = = = 1+ − M 2 (4.133) c γRT T 2 o s o r o  

T(K) 300 calorically perfect 250 ideal gas 200 R = 0.287 kJ/(kg K) γ = 7/5 150 = 300 K 100

50

M2 0 2 4 6 8 10

Figure 4.2: Static temperature versus Mach number squared

P(bar) 1

0.8 calorically perfect ideal gas 0.6 R = 0.287 kJ/(kg K) γ = 7/5 0.4 = 1 bar

0.2

M2 0 2 4 6 8 10

Figure 4.3: Static pressure versus Mach number squared

Example 4.4 Airplane problem

CC BY-NC-ND. 28 October 2019, J. M. Powers. 76 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

ρ(kg/m3)

1.2 calorically perfect 1 ideal gas

0.8 R = 0.287 kJ/(kg K) γ = 7/5 0.6 stagnation density = 1.16 kg/m3 0.4

0.2

M2 0 2 4 6 8 10

Figure 4.4: Static density versus Mach number squared

Given: An airplane is flying into still air at u = 200 m/s. The ambient air is at 288 K and 101.3 kPa. Find: Temperature, pressure, and density at nose of airplane Assume: Steady isentropic flow of CPIG Analysis: In the steady wave frame, the ambient conditions are static while the nose conditions are stagnation.

u u 200 m/s M = = = =0.588 (4.134) c √γRT 7 J 5 287 kgK 288 K r   so

1 T = T 1+ M 2 , (4.135) o 5   1 = (288 K) 1+ 0.5882 , (4.136) 5   = 307.9 K (4.137) 5 1 2 ρ = ρ 1+ M 2 (4.138) o 5   5 101.3 kPa 1 2 = 1+ 0.5882 , (4.139) 0.287 kJ 288 K 5 kgK   = 1.45 kg/m3 (4.140) 7 1 2 P = P 1+ M 2 , (4.141) o 5   7 1 2 = (101.3 kPa) 1+ 0.5882 (4.142) 5   = 128 kPa (4.143)

Note the temperature, pressure, and density all rise in the isentropic process. In this wave frame, the kinetic energy of the flow is being converted isentropically to thermal energy.

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.2. FLOW WITH AREA CHANGE 77

Example 4.5 Pressure measurement in compressible flows4

See Figure 4.5.

Air at 100 F z g Static State "1"

12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 8 inches 12345678901 Stagnation 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 State "o" 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 12345678901 123456789011234567890123456789012345678901212345678901234567890123456789012123456789012345678901234567890112345678901 1234567890112345678901234567890123456789012123456789012345678901234567890121234567890123456789012345678901 1234567890112345678901234567890123456789012123456789012345678901234567890121234567890123456789012345678901 Mercury 1234567890112345678901234567890123456789012123456789012345678901234567890121234567890123456789012345678901 1234567890112345678901234567890123456789012123456789012345678901234567890121234567890123456789012345678901 1234567890112345678901234567890123456789012123456789012345678901234567890121234567890123456789012345678901 1234567890112345678901234567890123456789012123456789012345678901234567890121234567890123456789012345678901 1234567890112345678901234567890123456789012123456789012345678901234567890121234567890123456789012345678901 1234567890112345678901234567890123456789012123456789012345678901234567890121234567890123456789012345678901 1234567890112345678901234567890123456789012123456789012345678901234567890121234567890123456789012345678901 1234567890112345678901234567890123456789012123456789012345678901234567890121234567890123456789012345678901

Figure 4.5: Compressible pitot tube sketch

ft Given: Air at u = 750 s , Mercury manometer which reads a change in height of 8 inches.

Find: Static pressure of air in psia

Assume: Ideal gas behavior for air

Analysis: First consider the manometer which is governed by fluid statics. In fluid statics, there is no motion, thus there are no viscous forces or fluid inertia; one thus has a balance between surface and body forces. Consider the linear momentum equation:

dv ρ = P + ρ g + τ (4.144) dt −∇ Hg ∇ · 0 = P + ρ g (4.145) −∇ Hg dP = ρ g (4.146) dz Hg z P P = ρ g (z z ) (4.147) 1 − o Hg z 1 − o lbm 1 lbf s2 ft 1 ft P P = 845.9 32.2 (0 in ( 8 in)) (4.148) 1 − o ft3 32.2 ft lbm − s2 − − 12 in         lbf P P = 563.9 (4.149) 1 − o − ft2 lbf P P = 563.9 (4.150) o − 1 ft2 lbf P = P + 563.9 (4.151) o 1 ft2

4adopted from White, 9.26, p. 584

CC BY-NC-ND. 28 October 2019, J. M. Powers. 78 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

Now calculate the local Mach number

T1 =100+460 = 560 R (4.152) u1 M1 = (4.153) √γRT1 ft 750 s M1 = (4.154) ft2 (1.4) 1, 717 s2 R (560 R) r   M1 = 0.646 (4.155)

Isentropic flow relations relate stagnation to static properties, so for air

1 3.5 P = P 1+ M 2 (4.156) o 1 5 1   1 3.5 P = P 1+ (0.646)2 (4.157) o 1 5   Po = 1.324P1 (4.158)

Substituting from the measured pressure difference

lbf P + 563.9 = 1.324P (4.159) 1 ft2 1 lbf 0.324P = 563.9 (4.160) − 1 − ft2 lbf 563.9 2 P = − ft (4.161) 1 0.324 − lbf P = 1, 740 (4.162) 1 ft2 lbf lbf P = (1.324) 1, 740 = 2, 304 (4.163) o ft2 ft2   lbf 1 ft 2 P = 1, 740 = 12.1 psia (4.164) 1 ft2 12 in     lbf 1 ft 2 P = 2, 304 = 16.0 psia (4.165) o ft2 12 in     What might one estimate if one did not account for compressibility effects? Assume one had the same static pressure and calculate what velocity one would predict.

First calculate the static density.

P1 ρ1 = (4.166) RT1 lbf 1, 740 ft2 ft lbm ρ1 = 32.2 (4.167) ft2 lbf s2 1, 717 s2 R (560 R)    lbm ρ = 0.05827 (4.168) 1 ft3

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.2. FLOW WITH AREA CHANGE 79

One would then use an incompressible Bernoulli equation:

ρ (0)2 ρ u2 P + = P + 1 1 (4.169) o 2 1 2

2 (Po P1) u1 = − (4.170) s ρ1

lbf 2 563.9 ft2 ft lbm u1 = v lbm 32.2 2 (4.171) u 0.05827 3 lbf s u ft   u t ft u = 789.4 (4.172) 1 s So the relative error in using the incompressible approximation would be 789.4 750 Error = − =5.3% (4.173) 750

Example 4.6 Adiabatic Duct Flow5

ft ◦ Given: Air flowing adiabatically through a duct. At section 1, u1 = 400 s ,T1 = 200 F, P1 = ft 35 psia. Downstream u2 =1, 100 s , P2 = 18 psia.

Find: M ,u , Po2 2 max Po1

Assume: Calorically perfect ideal gas, steady, one-dimensional flow

Analysis:

Some preliminaries:

T1 =200+460 = 660 R (4.174) Btu ft lbf lbm ft ft2 c = 0.240 779 32.17 = 6, 015 (4.175) p lbm R Btu lbf s2 s2 R       ft lbf lbm ft ft2 R = 53.34 32.17 = 1, 716 (4.176) lbm R lbf s2 s2 R     (4.177)

Energy conservation gives stagnation conditions at state 1

u2 h + 1 = h (4.178) 1 2 o1 5adopted from White’s 9.30, p. 585

CC BY-NC-ND. 28 October 2019, J. M. Powers. 80 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

u2 c T + 1 = c T (4.179) p 1 2 p o1 2 u1 To1 = T1 + (4.180) 2cp 2 ft 400 s To1 = 660 R + (4.181)  ft 2 2 6, 015 s2 R

To1 = 673 R   (4.182)

Note since in adiabatic flow ho is a constant, ho2 = ho1 and since ideal gas To2 = To1 So

To2 = 673 R (4.183) 2 u2 T2 = To2 (4.184) − 2cp 2 ft 1, 100 s T2 = 673 R (4.185) −  ft2 2 6, 015 s2 R

T2 = 572 R   (4.186)

Calculate the Mach numbers:

c1 = γRT1 (4.187) ft2 p ft c = 1.4 1, 716 (660 R) = 1, 259 (4.188) 1 s2 R s s   u 400 ft M = 1 = s = 0.318 (4.189) 1 c ft 1 1, 259 s

c2 = γRT2 (4.190) ft2 p ft c = 1.4 1, 716 (572 R) = 1, 173 (4.191) 2 s2 R s s   u 1, 100 ft M = 2 = s = 0.938 (4.192) 2 c ft 2 1, 173 s

Since for CPIG air one has

− 7 P 1 2 = 1+ M 2 (4.193) Po 5   7 1 2 P = P 1+ M 2 (4.194) o 5 7   1 2 P = (35 psia) 1+ 0.3182 = 37.54 psia (4.195) o1 5   7 1 2 P = (18 psia) 1+ 0.9382 = 31.74 psia (4.196) o2 5   P 31.74 psia o2 = = 0.845 (4.197) Po1 37.54 psia

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.2. FLOW WITH AREA CHANGE 81

Stagnation pressure drop indicates that friction was present. If one computed an entropy change one would see an increase in entropy.

The maximum velocity is found by converting all the thermal energy to kinetic energy. Taking zero thermal energy to correspond to absolute zero (despite the fact that air would not be a gas at this point) one could estimate u2 h = max (4.198) o 2 u2 c T = max (4.199) p o 2

umax = 2cpTo (4.200) ft2 p ft u = 2 6, 015 (673 R) = 2, 845 (4.201) max s2 R s s  

4.2.2 Sonic properties Let “*” denote a property at the sonic state M 2 1 ≡

1 T γ 1 2 − 2 ∗ = 1+ − 1 = (4.202) To 2 γ +1   1 1 γ−1 γ−1 ρ γ 1 2 − 2 ∗ = 1+ − 1 = (4.203) ρo 2 γ +1   γ   γ γ−1 γ−1 P γ 1 2 − 2 ∗ = 1+ − 1 = (4.204) P 2 γ +1 o     1/2 c γ 1 2 − 2 ∗ = 1+ − 1 = (4.205) c 2 γ +1 o   r 2γ u = c = γRT = RTo (4.206) ∗ ∗ ∗ γ +1 r p If air γ =7/5 and T ∗ = 0.8333 (4.207) To ρ ∗ = 0.6339 (4.208) ρo P ∗ = 0.5283 (4.209) Po c ∗ = 0.9129 (4.210) co

CC BY-NC-ND. 28 October 2019, J. M. Powers. 82 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

4.2.3 Effect of area change Influence of mass equation must be considered. So far only looked at energy has been examined. In the isentropic limit the mass, momentum, and energy equation for a CPIG reduce to dρ du dA + + = 0 (4.211) ρ u A ρudu + dP = 0 (4.212) dP dρ = γ (4.213) P ρ Substitute energy then mass into momentum: P ρudu + γ dρ = 0 (4.214) ρ P ρ ρ ρudu + γ du dA = 0 (4.215) ρ −u − A P 1 1  du + γ du dA = 0 (4.216) ρ −u2 − uA   γP/ρ P dA du 1 = γ (4.217) − u2 ρ uA   du γP/ρ γP/ρ dA 1 = (4.218) u − u2 u2 A   du 1 1 dA 1 = (4.219) u − M 2 M 2 A   du dA M 2 1 = (4.220) u − A du 1 dA  = (4.221) u M 2 1 A − Figure 4.6 gives show the performance of a fluid in a variable area duct. It is noted that equation singular when M 2 =1 • if M 2 = 1, one needs dA =0 • area minimum necessary to transition from subsonic to supersonic flow!! • can be shown area maximum not relevant • Consider A at a sonic state. From the mass equation:

ρuA = ρ u A (4.222) ∗ ∗ ∗

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.2. FLOW WITH AREA CHANGE 83

Consider u > 0

Subsonic Subsonic Diffuser Nozzle

2 dA > 0, M < 1 so 2 du < 0, flow slows down dA < 0, M <1 so dp > 0 du > 0, flow speeds up dp < 0

Supersonic Supersonic Nozzle Diffuser

2 2 dA > 0, M >1 so dA < 0, M > 1 so du > 0, flow speeds up du < 0, flow slows down dp < 0 dp > 0

Figure 4.6: Behavior of fluid in sub- and supersonic nozzles and diffusers

ρuA = ρ c A (4.223) ∗ ∗ ∗ A ρ 1 = ∗ c , (4.224) A ρ ∗ u ∗ ρ 1 = ∗ γRT , (4.225) ρ ∗ u ρ p√γRT √γRT = ∗ ∗ (4.226) ρ √γRT u ρ T 1 = ∗ ∗ , (4.227) ρ r T M ρ ρo T To 1 = ∗ ∗ (4.228) ρ ρ T T M o r o Substitute from earlier-developed relations and get:

1 γ+1 A 1 2 γ 1 2 γ−1 = 1+ − M 2 (4.229) A M γ +1 2 ∗    Figure 4.7 shows the performance of a fluid in a variable area duct.

Note:

CC BY-NC-ND. 28 October 2019, J. M. Powers. 84 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

A/A* 6

5

4 calorically perfect ideal gas 3 R = 0.287 kJ/(kg K)

2 γ = 7/5

1

M 0 0.5 1 1.5 2 2.5 3

Figure 4.7: Area versus Mach number for a calorically perfect ideal gas

A has a minimum value of 1 at M =1 • A∗ A For each A > 1, there exist two values of M • ∗ A as M 0 or M • A∗ →∞ → →∞ 4.2.4 Choking Consider mass flow rate variation with pressure difference

small pressure difference gives small velocity, small mass flow • as pressure difference grows, velocity and mass flow rate grow • velocity is limited to sonic at a particular duct location • this provides fundamental restriction on mass flow rate • can be proven rigorously that sonic condition gives maximum mass flow rate •

m˙ max = ρ u A (4.230) ∗ ∗ ∗ 1 2 γ−1 2γ if ideal gas: = ρo RTo A (4.231) γ +1 γ +1 ∗   r  1 1/2 2 γ−1 2 = ρo γRToA (4.232) γ +1 γ +1 ∗     1 γ+1 p 2 2 γ−1 = ρo γRToA (4.233) γ +1 ∗   p A flow which has a maximum mass flow rate is known as choked flow. Flows will choke at area minima in a duct.

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.2. FLOW WITH AREA CHANGE 85

Example 4.7 Isentropic area change problem with choking6

Given: Air with stagnation conditions Po = 200 kPa To = 500 K flows through a throat to an exit Mach number of 2.5. The desired mass flow is 3.0 kg/s,

Find: a) throat area, b) exit pressure, c) exit temperature, d) exit velocity, and e) exit area.

Assume: CPIG, isentropic flow, γ =7/5

Analysis: Po 200 kPa 3 ρo = = =1.394 kg/m (4.234) RTo (0.287 kJ/kg)(500 K) Since it necessarily flows through a sonic throat:

1 γ+1 2 2 γ−1 m˙ = ρ γRT A∗ (4.235) max o γ +1 o   m˙ max p A∗ = 1 γ+1 (4.236) − 2 2 γ 1 ρo γ+1 √γRTo   3 kg/s = , (4.237) kg J 1.394 m3 (0.5787) 1.4 287 kg K (500 K)   r   = 0.008297 m2 (4.238)

Since Me is known, use the isentropic relations to find other exit conditions. − γ γ 1 γ−1 P = P 1+ − M 2 , (4.239) e o 2 e   − 1 3.5 = (200 kPa) 1+ 2.52 , (4.240) 5   = 11.71 kPa (4.241) − γ 1 1 T = T 1+ − M 2 , (4.242) e o 2 e   − 1 1 = (500 K) 1+ 2.52 , (4.243) 5   = 222.2 K (4.244) Note

Pe ρe = , (4.245) RTe 11.71 kPa = , (4.246) kJ 0.287 kgK (222.2 K)  kg  = 0.1834 (4.247) m3 6adopted from White, Fluid Mechanics McGraw-Hill: New York, 1986, p. 529, Ex. 9.5

CC BY-NC-ND. 28 October 2019, J. M. Powers. 86 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

Now the exit velocity is simply

J m u = M c = M γRT =2.5 1.4 287 (222.2 K) = 747.0 (4.248) e e e e e kg K s s   p Now determine the exit area:

1 γ+1 A∗ 2 γ 1 2 γ−1 A = 1+ − M 2 (4.249) M γ +1 2 e e    0.008297 m2 5 1 3 = 1+ 2.52 , (4.250) 2.5 6 5    = 70.0219 m2 (4.251)

Example 4.8 Discharge Problem7

◦ 3 Given: Air in tank, Po = 700 kPa, To = 20 C, V = 1.5 m . Throat area in converging nozzle of 0.65 cm2, exhausting to 1 atm environment

Find: Time for pressure in tank to decrease to 500 kPa.

Assume: CPIG, stagnation temperature constant (so small heat transfer to tank in time of opera- tion)

Analysis: First, To = 20+273 = 293 K

Now check for choked flow! At the initial state P 101.3 kPa atm = =0.145 (4.252) Po 700 kPa But for air P∗ = 0.5283, so the flow must be choked at the exit and the mass flow is restricted. Po (Further expansion takes place outside the nozzle)

For choked flow one has

1 γ+1 2 2 γ−1 m˙ = ρ γRT A∗ (4.253) e o γ +1 o   p1 γ+1 2 γ−1 Po 2 = γRT A∗ (4.254) RT γ +1 o  o    p P J 1 m 2 = o 0.5787 1.4 287 (293 K) 0.65 cm2 (4.255)  J  kg K 100 cm 287 (293 K) s     kg K  = 1.5348 10−7P  (4.256) × o 7from White, 9.33,35

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.3. NORMAL SHOCK WAVES 87

Now mass conservation gives d m = m˙ (4.257) dt cv − e d (ρ V ) = m˙ (4.258) dt o − e d P o V = m˙ (4.259) dt RT − e  o  dP RT o = o m˙ (4.260) dt − V e J dP 287 kg K (293 K) o = 1.5348 10−7P (4.261) dt − 1.5 m 3 × o dP o = 0.008604P (4.262) dt − o P = A exp( 0.008604t) (4.263) o −

Use initial value of Po to fix the constant A so

P = 700exp( 0.008604t) (4.264) o −

When does Po = 500 kPa? 500 500 = 700exp( 0.008604t)ln = 0.008604t (4.265) − 700 − 1 500 t = ln = 39.1 s (4.266) −0.008604 700

4.3 Normal shock waves

This section will develop relations for normal shock waves in fluids with general equations of state. It will be specialized to calorically perfect ideal gases to illustrate the general features of the waves.

Assumptions for this section

one-dimensional flow • steady flow • no area change • viscous effects and wall friction do not have time to influence flow • heat conduction and wall heat transfer do not have time to influence flow •

CC BY-NC-ND. 28 October 2019, J. M. Powers. 88 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

vp = v2 D u2 u = - D v = v2 v = 0 P2 P1 P2 P1 ρ2 ρ1 ρ2 ρ1

Laboratory Frame Steady Frame x* x u = v - D; v = u + D x = x* - D t; x* = x + D t

Figure 4.8: Normal shock sketch

The piston problem as sketched in Figure 4.8 will be considered. Physical problem: Drive piston with known velocity v into fluid at rest (v = 0) with known properties, • p 1 P1, ρ1 in the x∗ laboratory frame Determine disturbance speed D • Determine disturbance properties v , P , ρ • 2 2 2 in this frame of reference unsteady problem • Transformed Problem:

use Galilean transformation x = x∗ Dt, u = v D to transform to the frame in which • − − the wave is at rest, therefore rending the problem steady in this frame

solve as though D is known to get downstream “2” conditions: u (D),P (D), ... • 2 2 invert to solve for D as function of u , the transformed piston velocity: D(u ) • 2 2 back transform to get all variables as function of v , the laboratory piston velocity: • 2 D(v2),P2(v2), ρ2(v2), ...

4.3.1 Governing equations Under these assumptions the conservation principles in conservative form and equation of state are in the steady frame as follows:

d (ρu) = 0 (4.267) dx d ρu2 + P = 0 (4.268) dx  CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.3. NORMAL SHOCK WAVES 89

d u2 ρu h + = 0 (4.269) dx 2    h = h(P, ρ) (4.270)

Upstream conditions are ρ = ρ , P = P , u = D. With knowledge of the equation of 1 1 − state, one gets h = h1. Integrating the equations from upstream to state “2” gives:

ρ2u2 = ρ1D (4.271) 2 − 2 ρ2u2 + P2 = ρ1D + P1 (4.272) u2 D2 h + 2 = h + (4.273) 2 2 1 2 h2 = h(P2, ρ2) (4.274)

4.3.2 Rayleigh line Work on the momentum equation:

2 2 P2 = P1 + ρ1D ρ2u2 (4.275) 2 2 − 2 2 ρ1D ρ2u2 P2 = P1 + (4.276) ρ1 − ρ2

2 2 2 2 1 Since mass gives ρ2u2 = ρ1D one gets an equation for the Rayleigh Line, a line in (P, ρ ) space:

1 1 P = P + ρ2D2 (4.277) 2 1 1 ρ − ρ  1 2  Note:

Rayleigh line passes through ambient state • Rayleigh line has negative slope • magnitude of slope proportional to square of wave speed • independent of state and energy equations •

4.3.3 Hugoniot curve Operate on the energy equation, using both mass and momentum to eliminate velocity. First eliminate u2 via the mass equation:

CC BY-NC-ND. 28 October 2019, J. M. Powers. 90 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

u2 D2 h + 2 = h + (4.278) 2 2 1 2 1 ρ D 2 D2 h + 1 = h + (4.279) 2 2 ρ 1 2  2  D2 ρ 2 h h + 1 1 = 0 (4.280) 2 − 1 2 ρ −  2  ! D2 ρ2 ρ2 h h + 1 − 2 = 0 (4.281) 2 − 1 2 ρ2  2  D2 (ρ ρ )(ρ + ρ ) h h + 1 − 2 1 2 = 0 (4.282) 2 − 1 2 ρ2  2  Now use the Rayleigh line to eliminate D2:

1 2 1 1 1 − D = (P2 P1) 2 (4.283) − ρ1 ρ1 − ρ2     1 1 ρ ρ − D2 = (P P ) 2 − 1 (4.284) 2 − 1 ρ2 ρ ρ  1   1 2  1 ρ ρ D2 = (P P ) 1 2 (4.285) 2 − 1 ρ2 ρ ρ  1   2 − 1  so the energy equation becomes 1 1 ρ ρ (ρ ρ )(ρ + ρ ) h h + (P P ) 1 2 1 − 2 1 2 = 0 (4.286) 2 − 1 2 2 − 1 ρ2 ρ ρ ρ2  1   2 − 1   2  1 1 ρ + ρ h h (P P ) 1 2 = 0 (4.287) 2 − 1 − 2 2 − 1 ρ ρ  1   2  1 1 1 h h (P P ) + = 0 (4.288) 2 − 1 − 2 2 − 1 ρ ρ  2 1  (4.289) Solving finally for the enthalpy difference, one finds 1 1 1 h h = (P P ) + (4.290) 2 − 1 2 − 1 2 ρ ρ    2 1  This equation is the Hugoniot equation. enthalpy change equals pressure difference times mean volume • independent of wave speed D and velocity u • 2 independent of equation of state •

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.3. NORMAL SHOCK WAVES 91

4.3.4 Solution procedure for general equations of state The shocked state can be determined by the following procedure:

specify and equation of state h(P, ρ) • substitute the equation of state into the Hugoniot to get a second relation between P • 2 and ρ2. use the Rayleigh line to eliminate P in the Hugoniot so that the Hugoniot is a single • 2 equation in ρ2 solve for ρ as functions of “1” and D • 2 back substitute to solve for P , u , h , T as functions of “1” and D • 2 2 2 2 invert to find D as function of “1” state and u • 2 back transform to laboratory frame to get D as function of “1” state and piston velocity • v2 = vp

4.3.5 Calorically perfect ideal gas solutions Follow this procedure for the special case of a calorically perfect ideal gas.

h = c (T T )+ h (4.291) p − o o P = ρRT (4.292) so P P h = c o + h (4.293) p Rρ − Rρ o  o  c P P h = p o + h (4.294) R ρ − ρ o  o  c P P h = p o + h (4.295) c c ρ − ρ o p − v  o  γ P P h = o + h (4.296) γ 1 ρ − ρ o −  o  Evaluate at states 1 and 2 and substitute into Hugoniot: γ P P γ P P 2 o + h 1 o + h γ 1 ρ − ρ o − γ 1 ρ − ρ o  −  2 o    −  1 o   1 1 1 = (P P ) + 2 − 1 2 ρ ρ    2 1  CC BY-NC-ND. 28 October 2019, J. M. Powers. 92 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

γ P P 1 1 1 2 1 (P P ) + = 0 γ 1 ρ − ρ − 2 − 1 2 ρ ρ −  2 1     2 1  γ 1 1 1 γ 1 1 1 P P = 0 2 γ 1 ρ − 2ρ − 2ρ − 1 γ 1 ρ − 2ρ − 2ρ  − 2 2 1   − 1 2 1  γ +1 1 1 γ +1 1 1 P P = 0 2 2(γ 1) ρ − 2ρ − 1 2(γ 1) ρ − 2ρ  − 2 1   − 1 2  γ +1 1 1 γ +1 1 1 P P = 0 2 γ 1 ρ − ρ − 1 γ 1 ρ − ρ  2 1   1 2  − − γ+1 1 1 γ 1 ρ1 ρ2 − − P2 = P1 γ+1 1 1 γ 1 ρ2 ρ1 − − a hyperbola in (P, 1 ) space • ρ 1 γ 1 1 − causes P2 , note for γ =1.4, ρ2 6ρ1 for infinite pressure • ρ2 → γ+1 ρ1 →∞ → 1 γ 1 as ,P2 P1 − , note negative pressure, not physical here • ρ2 →∞ → − γ+1 The Rayleigh line and Hugoniot curves are sketched in Figure 4.9.

P (kPa)

500 shocked state P 2 400 excluded zone slope of Rayleigh line < 0 excluded 300 zone, Rayleigh line, slope ~ D 2 1/ ρ < 1/ρmin from mass and momentum

200 Hugoniot, initial state from energy 100 P 1 excluded zone, 2nd law violation

1/ ρ 1/ ρ 2 3 4 5 6 7 1/ ρ (kg/m3 ) 2 1 -( γ-1) P1 γ+1 excluded zone, negative pressure

1/ ρ = (γ-1) 1 min (γ+1) ρ1

Figure 4.9: Rayleigh line and Hugoniot curve.

Note: intersections of the two curves are solutions to the equations • the ambient state “1” is one solution •

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.3. NORMAL SHOCK WAVES 93

the other solution “2” is known as the shock solution • shock solution has higher pressure and higher density • higher wave speed implies higher pressure and higher density • a minimum wavespeed exists • – occurs when Rayleigh line tangent to Hugoniot – occurs for very small pressure changes – corresponds to a sonic wave speed – disturbances are acoustic

if pressure increases, can be shown entropy increases • if pressure decreases (wave speed less than sonic), entropy decreases; this is non- • physical

Substitute Rayleigh line into Hugoniot to get single equation for ρ2

γ+1 1 1 2 2 1 1 γ 1 ρ1 ρ2 − − P1 + ρ1D = P1 γ+1 1 1 (4.297) ρ1 − ρ2 γ 1 ρ2 ρ1   − − This equation is quadratic in 1 and factorizable. Use computer algebra to solve and get ρ2 two solutions, one ambient 1 = 1 and one shocked solution: ρ2 ρ1

1 1 γ 1 2γ P = − 1+ 1 (4.298) ρ ρ γ +1 (γ 1) D2 ρ 2 1  − 1  The shocked density ρ2 is plotted against wave speed D for CPIG air in Figure 4.10.

Note density solution allows allows all wave speeds 0

CC BY-NC-ND. 28 October 2019, J. M. Powers. 94 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

strong shock 3 ρ2 (kg/m ) limit 7 6 calorically perfect 5 ideal air 4 exact γ = 7/5

3 solution R = 0.287 kJ/(kg K)

2 1 D (m/s) 500 1000 1500 2000 2500 3000

D = Dmin = c1

Figure 4.10: Shock density vs. shock wave speed for calorically perfect ideal air.

Back substitute into Rayleigh line and mass conservation to solve for the shocked pressure and the fluid velocity in the shocked wave frame: 2 γ 1 P = ρ D2 − P (4.299) 2 γ +1 1 − γ +1 1 γ 1 2γ P u = D − 1+ 1 (4.300) 2 − γ +1 (γ 1) D2 ρ  − 1 

The shocked pressure P2 is plotted against wave speed D for CPIG air in Figure 4.11 including both the exact solution and the solution in the strong shock limit. Note for these parameters, the results are indistinguishable.

P2 (Pa)

6 8. x 10 calorically perfect 6 6. x 10 ideal air ambient = γ = 7/5 6 4. x 10 100000 Pa R = 0.287 kJ/(kg K) exact 6 solution and 2. x 10 strong shock limit D (m/s) 500 1000 1500 2000 2500 3000

D = Dmin = c1

Figure 4.11: Shock pressure vs. shock wave speed for calorically perfect ideal air.

The shocked wave frame fluid particle velocity u2 is plotted against wave speed D for CPIG air in Figure 4.12.

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.3. NORMAL SHOCK WAVES 95

u2 (m/s) D (m/s) 500 1000 1500 2000 2500 3000 -100 calorically perfect -200 strong shock ideal air γ = 7/5 -300 limit R = 0.287 kJ/(kg K) -400 exact -500 solution

u1 = - c1 D = Dmin = c1

Figure 4.12: Shock wave frame fluid particle velocity vs. shock wave speed for calorically perfect ideal air.

2 The shocked wave frame fluid particle velocity M 2 = ρ2u2 is plotted against wave speed 2 γP2 D for CPIG air in Figure 4.13.

2 M2 calorically perfect 1 ideal air γ = 7/5 0.8 exact R = 0.287 kJ/(kg K) solution 0.6 strong shock 0.4 limit 0.2

D (m/s) 0 500 1000 1500 2000 2500 3000

D = Dmin = c1 2 M2 = 1

Figure 4.13: Mach number squared of shocked fluid particle vs. shock wave speed for calor- ically perfect ideal air.

2 Exercise: For the conditions shown in the plot of M2 vs. D do the detailed calculations to demonstrate the plot is correct.

Note in the steady frame that The Mach number of the undisturbed flow is (and must be) > 1: supersonic • The Mach number of the shocked flow is (and must be) < 1: subsonic •

CC BY-NC-ND. 28 October 2019, J. M. Powers. 96 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

Transform back to the laboratory frame u = v D: − γ 1 2γ P v D = D − 1+ 1 (4.301) 2 − − γ +1 (γ 1) D2 ρ  − 1  γ 1 2γ P v = D D − 1+ 1 (4.302) 2 − γ +1 (γ 1) D2 ρ  − 1  Manipulate the above equation and solve the resulting quadratic equation for D and get

γ +1 γP γ +1 2 D = v 1 + v2 (4.303) 4 2 ± ρ 2 4 s 1  

Now if v2 > 0, one expects D > 0 so take positive root, also set velocity equal piston velocity v2 = vp

γ +1 γP γ +1 2 D = v + 1 + v2 (4.304) 4 p ρ p 4 s 1   Note:

acoustic limit: as v 0, D c ; the shock speed approaches the sound speed • p → → 1 strong shock limit: as v ,D γ+1 v • p →∞ → 2 p

The shock speed D is plotted against piston velocity vp for CPIG air in Figure 4.14. Both the exact solution and strong shock limit are shown.

calorically perfect D (m/s) ideal air exact γ = 7/5 1200 solution R = 0.287 kJ/(kg K) 1000 800 strong shock 600 acoustic limit limit, 400 D c1 200 v (m/s) 200 400 600 800 1000 p

Figure 4.14: Shock speed vs. piston velocity for calorically perfect ideal air.

If the Mach number of the shock is defined as

D Ms (4.305) ≡ c1

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.3. NORMAL SHOCK WAVES 97 one gets

γ +1 v v2 γ +1 2 M = p + 1+ p (4.306) s 4 √γRT γRT 4 1 s 1  

The shock Mach number Ms is plotted against piston velocity vp for CPIG air in Figure 4.15. Both the exact solution and strong shock limit are shown.

M s exact calorically perfect

3.5 solution ideal air 3 γ = 7/5 2.5 R = 0.287 kJ/(kg K)

2 strong shock 1.5 acoustic limit limit, 1 Ms 1 0.5 v (m/s) 200 400 600 800 1000 p

Figure 4.15: Shock Mach number vs. piston velocity for calorically perfect ideal air

Example 4.9 Normal shock problem8

Given: Air flowing through normal shock. Upstream u1 = 600 m/s, To1 = 500 K, Po1 = 700 kPa.

Find: Downstream conditions M ,u ,T , P , P and s s . 2 2 2 2 o2 2 − 1 Assume: calorically perfect ideal gas

Analysis: First get all local unshocked conditions.

2 u1 To1 = T1 + (4.307) 2cp 2 u1 T1 = To1 (4.308) − 2cp m 2 600 s T1 = 500 K , (4.309) − J 2 1004 .5 kg K = 320.81 K   (4.310)

c1 = γRT1, (4.311) p J = 1.4 287 (320.81 K), (4.312) kg K s   m = 359.0 (4.313) s 8adopted from White’s 9.46, p. 586

CC BY-NC-ND. 28 October 2019, J. M. Powers. 98 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

u1 M1 = , (4.314) c1 m 600 s = m , (4.315) 359.0 s = 1.671 (4.316) − 1 3.5 P = P 1+ M 2 , (4.317) 1 o1 5 1   − 1 3.5 = (700 kPa) 1+ (1.671)2 , (4.318) 5   = 148.1 kPa (4.319)

P1 ρ1 = , (4.320) RT1 148.1 kPa = , (4.321) kJ 0.287 kg K (320.81 K)  kg  = 1.609 (4.322) m3 1 γ+1 2 γ−1 A1 1 2 γ 1 2 = 1+ − M1 (4.323) A ∗ M γ +1 2 1 1    1 1.4+1 1 2 1.4 1 2 1.4−1 = 1+ − 1.6712 , (4.324) 1.671 1.4+1 2    = 1.311 (4.325)

m Now in this case it is fortunate because the incoming velocity D = 600 s is known. Note that the shock density only depends on D2, so one can be a little sloppy here with sign. Solve for the shocked state:

1 1 γ 1 2γ P = − 1+ 1 (4.326) ρ ρ γ +1 (γ 1) D2 ρ 2 1  − 1  1 1 1.4 1 2 (1.4) 148, 100 Pa = − 1+ (4.327) kg m 2 kg ρ2 1.609 3 1.4+1 (1.4 1) 600 1.609 3 ! m − s m m3 = 0.2890  (4.328) kg 1 ρ2 = m3 , (4.329) 0.2890 kg kg = 3.461 (4.330) m3

Now a variety of equations can be used to determine the remaining state variables. Mass gives u2:

ρ2u2 = ρ1u1 (4.331) ρ1u1 u2 = , (4.332) ρ2 kg m 1.609 m3 600 s = , (4.333)   kg 3.461 m 3 

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.3. NORMAL SHOCK WAVES 99

m = 278.9 . (4.334) s

Momentum gives P2 2 2 P2 + ρ2u2 = P1 + ρ1u1 (4.335) P = P + ρ u2 ρ u2 (4.336) 2 1 1 1 − 2 2 kg m 2 kg m 2 P = 148, 100 Pa + 1.609 600 3.461 278.9 (4.337) 2 m3 s − m3 s         P2 = 458, 125 Pa = 458 kPa (4.338) Remaining assorted variables are straightforward:

P2 T2 = (4.339) ρ2R 458, 125 Pa = , (4.340) kg J 3.461 m3 287 kg K = 461 .2 K   (4.341)

c2 = γRT2, (4.342) p J = 1.4 287 (461.2 K), (4.343) kg K s   m = 430.5 (4.344) s u2 M2 = , (4.345) c2 m 278.9 s = m , (4.346) 430.5 s = 0.648 (4.347) 1 T = T 1+ M 2 , (4.348) o2 2 5 2   1 = 461.2 K 1+ 0.6482 (4.349) 5   = 500 K unchanged as required (4.350) 1 3.5 P = P 1+ M 2 , (4.351) o2 2 5 2   1 3.5 = 458 kPa 1+ 0.6482 (4.352) 5   = 607.4 kPa dropped from unshocked state (4.353) T P s s = c ln 2 R ln 2 (4.354) 2 − 1 p T − P  1   1  J 461.2 K J 458 kPa = 1004.5 ln 287 ln (4.355) kg K 320.81 K − kg K 148.1 kPa     = 364.6 324.0, (4.356) − J = 40.6 (4.357) kg K 1 γ+1 2 γ−1 A2 1 2 γ 1 2 = 1+ − M2 (4.358) A ∗ M γ +1 2 2 2    CC BY-NC-ND. 28 October 2019, J. M. Powers. 100 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

1 1.4+1 1 2 1.4 1 2 1.4−1 = 1+ − 0.6482 , (4.359) 0.648 1.4+1 2    = 1.966. (4.360)

Since A2 = A1 = A,

A A2 A1∗ 1.311 = A = =0.667 (4.361) A2∗ 1.966 A2∗ Note the entropy increased despite not including any entropy-generating mechanisms in this model. Why? First, the differential equations themselves required the assumption of continuous differentiable functions. Our shock violates this. When one returns to the more fundamental control volume forms, it can be shown that the entropy-generating mechanism returns. From a continuum point of view, one can also show that the neglected terms, that momentum and energy diffusion, actually give rise to a smeared shock. These mechanisms generate just enough entropy to satisfy the entropy jump which was just calculated. Just as with Burger’s equation and the kinematic wave equation, the jumps are the same, diffusion simply gives a wave thickness.

Example 4.10 Piston Problem

m Given: A piston moving at vp = 1, 000 s is driven into Helium which is at rest in the ambient state at a pressure of P1 = 10 kPa, T1 = 50 K.

Find: The shock speed and post shock state.

Assume: Helium is calorically perfect and ideal

Analysis: For Helium,

γ = 1.667 (4.362) J R = 2077 (4.363) kg K γR c = , (4.364) p γ 1 − J 1.667 2, 077 kg K = , (4.365) 1.667 1  J− = 5, 192.5 . (4.366) kg K Ambient density

P1 ρ1 = , (4.367) RT1 10, 000 Pa = , (4.368) J 2, 077 kg K (50 K)   CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.3. NORMAL SHOCK WAVES 101

kg = 0.0963 (4.369) m3

c1 = γRT1, (4.370) p J = 1.667 2, 077 (50 K), (4.371) kg K s   m = 416.0 (4.372) s Now the wave speed D one gets from

2 γ +1 γP1 2 γ +1 D = vp + + vp (4.373) 4 s ρ1 4   1.667+1 m 1.667 (10, 000 Pa) m 2 1.667+1 2 = 1, 000 + + 1, 000 (4.374) 4 s v kg s 4 u 0.0963 m3   u     = 666.7+785.8, t (4.375) m = 1, 452.5 (4.376) s Strong shock limit is appropriate here as a quick check: γ +1 1.667+1 m m D v = 1, 000 =1, 333.3 (4.377) ∼ 2 p 2 s s   2 γ 1 P = ρ D2 − P (4.378) 2 γ +1 1 − γ +1 1 2 kg m 2 1.667 1 = 0.0963 1, 452.5 − (10, 000 Pa) (4.379) 1.667+1 m3 s − 1.667+1     = 152, 377 2, 500, (4.380) − = 149, 877 Pa = 150 kPa (4.381)

ρ2u2 = ρ1u1 (4.382) ρ (v D) = ρ (v D) (4.383) 2 2 − 1 1 − ρ (v D) = ρ (0 D) (4.384) 2 p − 1 − ρ D ρ = − 1 (4.385) 2 v D p − kg m 0.0963 m3 1, 452.5 s = − , (4.386) 1, 000 m 1, 452.5 m  s − s kg = 0.309 (4.387) m3 P2 T2 = (4.388) ρ2R 149, 877 Pa = , (4.389) kg J 0.309 m3 2, 077 kg K = 233 .5 K   (4.390)

CC BY-NC-ND. 28 October 2019, J. M. Powers. 102 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

4.3.6 Acoustic limit Consider that state 2 is a small perturbation of state 1 so that

ρ2 = ρ1 + ∆ρ (4.391)

u2 = u1 + ∆u1 (4.392)

P2 = P1 + ∆P (4.393) Substituting into the normal shock equations, one gets

(ρ1 + ∆ρ)(u1 + ∆u) = ρ1u1 (4.394) 2 2 (ρ1 + ∆ρ)(u1 + ∆u) +(P1 + ∆P ) = ρ1u1 + P1 (4.395) γ P + ∆P 1 γ P 1 1 + (u + ∆u)2 = 1 + u 2 (4.396) γ 1 ρ + ∆ρ 2 1 γ 1 ρ 2 1 − 1 − 1 Expanding, one gets

ρ1u1 +˜u1 (∆ρ)+ ρ1 (∆u)+(∆ρ) (∆u) = ρ1u1 2 2 2 2 ρ1u1 +2ρ1u1 (∆u)+ u1 (∆ρ)+ ρ1 (∆u) +2u1 (∆u) (∆ρ)+(∆ρ) (∆u) 2 +(P1 + ∆P) = ρ1u1 + P1 γ P 1 P 1 1 + ∆P 1 ∆ρ + ... + u 2 +2u (∆u)+(∆u)2 γ 1 ρ ρ − ρ2 2 1 1 −  1 1 1   γ P 1 = 1 + u 2 γ 1 ρ 2 1 − 1 Subtracting the base state and eliminating products of small quantities yields

u1 (∆ρ)+ ρ1 (∆u) = 0 (4.397) 2 2ρ1u1 (∆u)+ u1 (∆ρ) + ∆P = 0 (4.398) γ 1 P ∆P 1 ∆ρ + u (∆u) = 0 (4.399) γ 1 ρ − ρ2 1 −  1 1  In matrix form this is

u1 ρ1 0 ∆ρ 0 2 u1 2ρ1u1 1 ∆u = 0 (4.400)  γ P1 γ 1      2 u1 ∆P 0 − γ 1 ρ1 γ 1 ρ1  − −      As the right hand side is zero, the determinant must be zero and there must be a linear dependency of the solution. First check the determinant:

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.3. NORMAL SHOCK WAVES 103

2γ γ u 2 γ P u u u ρ 1 + 1 = 0 (4.401) 1 γ 1 1 − 1 − 1 γ 1 ρ γ 1 ρ2  −   − 1 − 1  u 2 1 P 1 (2γ (γ 1)) γ u 2 + γ 1 = 0 (4.402) γ 1 − − − γ 1 1 ρ − −  1  P u 2 (γ + 1) γ u 2 + γ 1 = 0 (4.403) 1 − 1 ρ  1  2 P1 2 u1 = γ = c1 (4.404) ρ1

So the velocity is necessarily sonic for a small disturbance!

Take ∆u to be known and solve a resulting 2 2 system: ×

u1 0 ∆ρ ρ1∆u γ P1 γ 1 = − (4.405) 2 γ 1 ρ γ 1 ρ1 ∆P u1∆u  − − 1 −     −  Solving yields

ρ ∆u ∆ρ = 1 (4.406) − γ P1 ρ1 q P1 ∆P = ρ1 γ ∆u (4.407) − s ρ1

4.3.7 Non-ideal gas solutions Non-ideal effects are important

near the critical point • for strong shocks • Some other points:

qualitative trends the same as for ideal gases • analysis is much more algebraically complicated • extraneous solutions often arise which must be discarded •

CC BY-NC-ND. 28 October 2019, J. M. Powers. 104 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

Example 4.11 Shock in van der Waals gas

m Given: Shock wave D = 500 s propagating into N2 at rest at T1 = 125 K, P1 =2 MPa.

Find: Shocked state

Assume: van der Waals equation of state accurately models gas behavior, specific heat constant.

Analysis:

First, some data for N2 are needed. At P1 = 2 MPa, N2 has a boiling point of 115.5 K, so J J the material is in the gas phase but very near the vapor dome. R = 296.8 kg K , cv = 744.8 kg K , Tc = 126.2 K, Pc =3, 390, 000 Pa.

Since the material is near the vapor dome, the van der Waals equation may give a good first correction for non-ideal effects. RT a P = (4.408) v b − v2 RT− P = aρ2 (4.409) 1 b − ρ − ρRT P = aρ2 (4.410) 1 bρ − − As derived earlier, the corresponding caloric equation of state is

T 1 1 e(T, v)= e + c (Tˆ)dTˆ + a (4.411) o v v − v ZTo  o 

Taking cv constant and exchanging v for ρ gives

e(T,ρ)= e + c (T T )+ a (ρ ρ) (4.412) o v − o o − Eliminating T in favor of P then gives

P + aρ2 (1 bρ) e(P,ρ)= eo + cv − To + a (ρo ρ) (4.413) ρR − ! −

and in terms of h = e + P/ρ:

P + aρ2 (1 bρ) P h(P,ρ)= eo + cv − To + a (ρo ρ)+ (4.414) ρR − ! − ρ and h h allows cancellation of the “o” state so that 2 − 1 2 2 P2 + aρ2 (1 bρ2) P1 + aρ1 (1 bρ1) P2 P1 h2 h1 = cv − − a (ρ2 ρ1)+ (4.415) − ρ2R − ρ1R ! − − ρ2 − ρ1

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.3. NORMAL SHOCK WAVES 105

The constants a and b are fixed so that an isotherm passing through the critical point, P = Pc,T = ∂P ∂2P 9 Tc, passes through with = 0 and 2 = 0 A standard analysis yields ∂v T ∂v T

27 R2T 2 a = c , (4.416) 64 Pc 2 J 2 27 296.8 kg K (126.2 K) = , (4.417) 64  3, 390,000 Pa Pa m6 = 174.6 (4.418) kg2 RT b = c , (4.419) 8Pc J 296.8 kg K (126.2 K) = , (4.420)  8 (3, 390,000 Pa) m3 = 0.00138 (4.421) kg Find the ambient density.

J 6 ρ1 296.8 kg K (125 K) Pa m 2, 000, 000 Pa = 174.6 ρ2 (4.422) 3 2 1 1 0.00138m ρ − kg − kg 1     Three solutions (from computer algebra): kg ρ = 69.0926 physical (4.423) 1 m3 kg ρ = (327.773+ 112.702 i) non-physical (4.424) 1 m3 kg ρ = (327.773+ 112.702 i) non-physical (4.425) 1 m3

kg Tabular data from experiments gives ρ1 = 71.28 m3 , error = (71.28 69.09)/71.28= 3%, so it seems the first root is the physical root. Note that the van der Waals prediction− is a significant improvement over P1 2,000,000 kg the which gives ρ = = = 53.91 3 , error = (71.28 53.91)/71.28= 21.4%! 1 RT1 296.8×125 m Even with this improvement there are much better (and more complicated!)− equations of state for materials near the vapor dome.

Now use the Rayleigh line and Hugoniot equations to solve for the shocked density: 1 1 P = P + ρ2D2 2 1 1 ρ − ρ  1 2  2 2 P2 + aρ2 (1 bρ2) P1 + aρ1 (1 bρ1) P2 P1 cv − − a (ρ2 ρ1)+ ρ2R − ρ1R ! − − ρ2 − ρ1 ! 1 1 1 (P P ) + = 0 − 2 2 − 1 ρ ρ    2 1  9Sonntag and Van Wylen, 1991, Introduction to Thermodynamics: Classical and Statistical, John Wiley: New York, p. 392.

CC BY-NC-ND. 28 October 2019, J. M. Powers. 106 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

Plugging in all the numbers into a computer algebra program yields the following solutions for ρ2: kg ρ = 195.309 shocked solution (4.426) 2 m3 kg ρ = 69.0926 inert solution (4.427) 2 m3 kg ρ = (85.74+ 657.9 i) non-physical solution (4.428) 2 m3 kg ρ = (85.74 657.9 i) non-physical solution (4.429) 2 − m3 The Rayleigh line then gives the pressure:

kg 2 m 2 1 1 P2 = 2, 000, 000 Pa + 69.0926 3 500 kg kg (4.430) m s 69.0926 3 − 195.309 3 !     m m P2 = 13, 162, 593 Pa = 13.2 MPa (4.431) The state equation gives the temperature.

2 P2 + aρ2 (1 bρ2) T2 = − (4.432) ρ2R 2 Pa m6 kg m3 kg 13, 162, 593 Pa + 174.6 2 195.3 3 1 0.00138 195.3 3 kg m − kg m =   (4.433)  kg  J    195.3 m3 296.8 kg K = 249.8 K    (4.434)

Note the temperature is still quite low relative to standard atmospheric conditions; it is unlikely at these low that any effects due to vibrational relaxation or dissociation will be important. Our assumption of constant specific heat is probably pretty good.

The mass equation gives the shocked particle velocity:

ρ2u2 = ρ1u1 (4.435) ρ1u1 u2 = (4.436) ρ2 kg m 69.0926 m3 500 s = (4.437)  kg 195.3 m3  m = 176.89 (4.438) s

An ideal gas approximation (γN2 =1.4) would have yielded 1 1 γ 1 2γ P = − 1+ 1 (4.439) ρ ρ γ +1 (γ 1) D2 ρ 2 1  − 1  1 1 1.4 1 2 (1.4) 2, 000, 000 Pa = − 1+ (4.440) kg m 2 kg ρ2 53.91 3 ! 1.4+1 (1.4 1) 500 53.91 3 ! m − s m kg ρ = 158.65 ideal gas approximation  (4.441) 2 m3 195.3 158.65 relative error = − = 18.8% (4.442) 195.3

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.4. FLOW WITH AREA CHANGE AND NORMAL SHOCKS 107

The Rayleigh line then gives the pressure:

kg 2 m 2 1 1 P2 = 2, 000, 000 Pa + 53.91 3 500 kg kg (4.443) m s 53.91 3 − 158.65 3 !     m m P2 = 10, 897, 783 Pa = 10.90 MPa (4.444) 13.2 10.9 relative error = − = 17.4% (4.445) 13.2

4.4 Flow with area change and normal shocks

This section will consider flow from a reservoir with the fluid at stagnation conditions to a constant pressure environment. The pressure of the environment is commonly known as the back pressure: Pb.

Generic problem: Given A(x), stagnation conditions and Pb, find the pressure, tempera- ture, density at all points in the duct and the mass flow rate.

4.4.1 Converging nozzle

A converging nozzle operating at several different values of Pb is sketched in Figure 4.16. The flow through the duct can be solved using the following procedure

check if Pb P • ≥ ∗ if so, set P = P • e b determine M from isentropic flow relations • e determine A from A relation • ∗ A∗ A A at any point in the flow where A is known, compute A∗ and then invert A∗ relation to • find local M Note: These flows are subsonic throughout and correspond to points a and b in Figure 4.16. •

If Pb = P then the flow is sonic at the exit and just choked. This corresponds to point ∗ • c in Figure 4.16.

If Pb

CC BY-NC-ND. 28 October 2019, J. M. Powers. 108 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

. . m/mmax e d c Pb 1

b Po Pe a

0 pb/p o p*/p o 1

P(x)/Po 1 a--subsonic exit b--subsonic exit P*/Po c--sonic exit d--choked, external expansion e--choked, external expansion

x xe

Figure 4.16: Converging nozzle sketch

4.4.2 Converging-diverging nozzle

A converging-diverging nozzle operating at several different values of Pb is sketched in Figure 4.17. The flow through the duct can be solved using the a very similar following procedure

set At = A • ∗ with this assumption, calculate Ae • A∗ determine M , M , both supersonic and subsonic, from A relation • esub esup A∗ determine P ,P , from M , M ; these are the supersonic and subsonic design • esub esup esub esup pressures

if P > P , the flow is subsonic throughout and the throat is not sonic. Use same • b esub procedure as for converging duct: Determine Me by setting Pe = Pb and using isentropic relations

if P >P >P , the procedure is complicated • esub b esup

– estimate the pressure with a normal shock at the end of the duct, Pesh – If P P , there is a normal shock inside the duct b ≥ esh – If Pb

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.4. FLOW WITH AREA CHANGE AND NORMAL SHOCKS 109

Pb

Po Pt Pe possible normal shock

P(x)/Po

1 a--subsonic exit b--subsonic exit c--subsonic design d--shock in duct P*/P o e-shock at end of duct sonic f--external compression throat g--supersonic design x h--external expansion xt xe . . m/mmax h g f e d c 1

b

a

0 P / P P*/Po 1 b o

Figure 4.17: Converging-diverging nozzle sketch

if P = P the flow is at supersonic design conditions and the flow is shockless • esup b if P

Example 4.12 Nozzle Problem10

2 2 Given: Air at To = 600 K flowing through converging-diverging nozzle. At =1 cm , Ae =3 cm , kg m˙ = 148.5 hr . Pitot tube at exit plane gives Poe = 200 kPa, Pe = 191.5 kPa.

10adopted from White’s 9.69, p. 588

CC BY-NC-ND. 28 October 2019, J. M. Powers. 110 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

Find: exit velocity, location of possible normal shock in duct, Mach number just upstream of normal shock

Assume: Air is calorically perfect and ideal

Analysis:

kg hr kg m˙ = 148.5 =0.04125 (4.446) hr 3600 s s Now if there is no shock, the stagnation pressure would be constant in the duct; one can use the choked flow formula to compare to the actual mass flow rate:

1 γ+1 2 γ−1 Po 2 m˙ = γRT A∗ (4.447) e RT γ +1 o o   1 1.4+1 p 2 200, 000 Pa 2 2 1.4−1 J 1 m = 1.4 287 (600 K) 1 cm2 (4.448) J 1.4+1 kg K 100 cm 287 (600 K)   s     kg K    kg = 200, 000 165 10−9 =0.033 (4.449) × × s  Now the actual mass flow is higher than this, so the stagnation pressure upstream must also be higher; therefore, there must be a shock in the duct which lowers the stagnation pressure. Use this equation to determine what the upstream stagnation pressure must be.

kg kg 1 0.04125 = P 165 10−9 (4.450) s o1 × × s Pa   Po1 = 250 kPa (4.451)

So P 200 kPa o2 = =0.800 (4.452) Po1 250 kPa The flow conditions could be deduced from this; one can also utilize the normal shock tables for air. These are valid only for a calorically perfect ideal air. Interpolating this table yields

M 1.83 (4.453) 1 ∼ M 0.61 (4.454) 2 ∼ The area ratio is determined from the isentropic flow tables. Recall that A∗ changes through a shock, so in this case one wants to use conditions upstream of the shock. From the tables at M1 =1.83 A1 one finds A∗ =1.4723 so,

A =1.4723 1 cm2 =1.4723 cm2 (4.455) 1 ×

Get the exit velocity. Even if there is a shock, the stagnation temperature is constant; thus, one has from energy conservation:

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.5. FLOW WITH FRICTION– 111

u2 h + e = h (4.456) e 2 o u = 2 (h h ) (4.457) e o − e = p2c (T T ) (4.458) p o − e q T = 2c T 1 e (4.459) p o − T s  o  γ−1 P γ = 2c T 1 e (4.460) v p o u − Poe ! u   t 1.4−1 J 191.5 kPa 1.4 = 2 1004.5 (600 K) 1 (4.461) v u kg K − 200 kPa ! u     t m = 121.9 (4.462) s

4.5 Flow with friction–Fanno flow

Wall friction is typically considered by modelling the wall shear as a constant. Wall friction is usually correlated with what is known as the Darcy friction factor: f, where

8τ f w (4.463) ≡ ρu2 Now in practice f is related to the local flow based on pipe diameter D: ReD

ρuD Re (4.464) D ≡ µ ǫ and roughness of the duct D , where ǫ is the average surface roughness.

ǫ f = f Re , (4.465) D D   For steady laminar duct flow, the friction factor is independent of ǫ. It turns out the Poiseuille flow solution gives the friction factor, which turns out to be 64 f = (4.466) ReD

CC BY-NC-ND. 28 October 2019, J. M. Powers. 112 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

If the flow is steady and turbulent, the friction factor is described by the following em- pirical formula known as the Colebrook equation:

1 ǫ/D 2.51 = 2.0 log + (4.467) f 1/2 − 10 3.7 Re f 1/2  D  Often one needs to iterate to find f for turbulent flows. Alternatively, one can use the Moody chart to estimate f. This is simply a graphical representation of the Colebrook formula. Most fluid texts will contain a Moody chart. While in principle f varies with a host of variables, in practice in a particular problem, it is often estimated as a constant. To get a grasp on the effects of wall friction, consider a special case of generalized one- dimensional flow: steady • one-dimensional • adiabatic • constant area duct • Darcy friction model • calorically perfect ideal gas • Our equations from the section on influence coefficients

2 dA (qw+τwu) ρu + τ + L dx w ρu ∂e dρ 1 − L ∂P ρ = | (4.468) dx A (u2 c2) − 2 dA (qw+τwu) c ρu uτ L dx w ρ ∂e du 1 − L − ∂P = |ρ (4.469) dx A ρ (u2 c2) − 2 2 dA 2 (qw+τwu) u c ρu + c τ + L dx w ρ ∂e dP 1 − L ∂P = |ρ (4.470) dx A (u2 c2) − reduce to 1 1+ ∂e dρ τ ρ ∂P = wL |ρ (4.471) dx A (u2 c2) − 1 1+ ∂e du uτ ρ ∂P = wL |ρ (4.472) dx − A ρ (u2 c2) − 2 u2 c + ∂e dP τ ρ ∂P = wL |ρ (4.473) dx A (u2 c2) − CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.5. FLOW WITH FRICTION–FANNO FLOW 113

Now for a circular duct

= 2πr (4.474) L A = πr2 (4.475) 2πr 2 4 L = = = (4.476) A πr2 r D For a calorically perfect ideal gas 1 P e = (4.477) γ 1 ρ − ∂e 1 1 = (4.478) ∂P γ 1 ρ ρ − ∂e 1 ρ = (4.479) ∂P γ 1 ρ − 1 = γ 1 (4.480) ∂e − ρ ∂P ρ 1 1+ ∂e = γ (4.481) ρ ∂P ρ

So making these substitutions yields dρ 4τ γ = w (4.482) dx D (u2 c2) − du 4uτ γ = w (4.483) dx − D ρ (u2 c2) − dP 4τ c2 + u2 (γ 1) = w − (4.484) dx D (u2 c2) −

Substituting for τw gives

dρ fρu2 γ = (4.485) dx 2D (u2 c2) − du fρu2u γ = (4.486) dx − 2D ρ (u2 c2) − dP fρu2 c2 + u2 (γ 1) = − (4.487) dx 2D (u2 c2) − Rearranging to place in terms of M 2 gives

dρ fρM 2 γ = (4.488) dx 2D (M 2 1) − CC BY-NC-ND. 28 October 2019, J. M. Powers. 114 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

du fM 2u γ = (4.489) dx − 2D (M 2 1) − dP fρu2 1+ M 2 (γ 1) = − (4.490) dx 2D (M 2 1) − Now with the definition of M 2 for the calorically perfect ideal gas, one gets

ρu2 M 2 = γP dM 2 u2 dρ 2ρu du ρu2 dP = + dx γP dx γP dx − γP 2 dx u2 fρM 2 γ 2ρu fM 2u γ ρu2 fρu2 1+ M 2 (γ 1) = + − γP 2D (M 2 1) γP − 2D (M 2 1) − γP 2 2D (M 2 1) −  −  − fM 4 γ 2fM 4 γ γfM 4 1+ M 2 (γ 1) = − 2D (M 2 1) − 2D (M 2 1) − 2D (M 2 1) − − − γfM 4 = 1 2 1 M 2 (γ 1) 2D (M 2 1) − − − − − γfM 4 γ 1  = 1+ M 2 − D (1 M 2) 2 −    So rearranging gives

(1 M 2) dM 2 dx − = f (4.491) 2 2 2 γ 1 D γ (M ) 1+ M −2 Integrate this expression from x = 0 to x = L where L is defined as the length at which ∗ ∗ the flow becomes sonic, so M 2 =1 at x = L . ∗

2 2 1 1 Mˆ dMˆ L∗ − dx 2 = f (4.492) 2   M ˆ 2 ˆ 2 γ 1 0 D Z γ M 1+ M −2 Z     An analytic solution for this integral is

1 M 2 1+ γ (1 + γ) M 2 fL − + ln = ∗ (4.493) γM 2 2γ 2+ M 2 (γ 1) D −

Example 4.13 Flow in a duct with friction11 11from White, 9.82, p. 589

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.5. FLOW WITH FRICTION–FANNO FLOW 115

ft Given: Air flowing in pipe D =1 in, L = 20 ft, P1 = 40 psia, u1 = 200 s , T1 = 520 R.

Find: Exit pressure P2,m ˙

Assume: calorically perfect ideal gas, Darcy friction factor models wall shear, constant viscosity

Analysis: First get the mass flow rate.

P1 ρ1 = (4.494) RT1 lbf in2 40 in2 144 ft2 ρ1 = (4.495)  ft lbf  53.34 lbm R (520 R)  lbm  ρ = 0.2077 (4.496) 1 ft3

m˙ = ρ1u1A1, (4.497) D 2 = ρ u π (4.498) 1 1 2   ! lbm ft 1 in 1 ft 2 = 0.2077 200 π (4.499) ft3 s 2 12 in     ! lbm = 0.2266 (4.500) s Now compute the friction factor. First for cast iron pipes, one has surface roughness ǫ =0.00085 ft, so

ǫ 0.00085 ft 12 in = =0.0102 (4.501) D 1 in 1 ft

−7 lbf s The Reynolds number is needed, which involves the viscosity. For air at 520 R, µ 4.08 10 ft2 so ∼ ×

lbm ft 1 0.2077 3 200 ft ρ1u1D ft s 12 ReD = = = 263, 739 (4.502) µ  −7 lbf s  lbm ft 4.08 10 2 32.17 2 × ft lbf s    Since ReD >> 2, 300, the flow is turbulent and one needs to use the Colebrook formula to estimate the Darcy friction factor:

1 ǫ/D 2.51 = 2.0 log + (4.503) f 1/2 − 10 3.7 Re f 1/2  D  0.0102 2.51 = 2.0 log + (4.504) − 10 3.7 263, 739f 1/2   Now reading the Moody chart gives f = 0.04. A numerical trial and error solution of the Colebrook equation gives

f =0.0384 (4.505)

CC BY-NC-ND. 28 October 2019, J. M. Powers. 116 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

Now find M1.

u1 M1 = , (4.506) √γRT1 200 ft = s , (4.507) ft lbf lbm ft 1.4 53.34 lbm R 32.17 lbf s2 (520 R) r = 0.1789   (4.508)

Now

fL ∗ 1 M 2 1+ γ (1 + γ) M 2 1 = − 1 + ln 1 (4.509) D γM 2 2γ 2+ M 2 (γ 1) 1 1 − 1 0.17892 1+1.4 (1+1.4)0.17892 = − + ln (4.510) 1.4 (0.1789)2 2 (1.4) 2+0.17892 (1.4 1) − = 18.804 (4.511) 1 18.804 12 ft L1∗ = , (4.512) 0.0384  = 40.81 ft (4.513)

so at a distance 40.81 ft from station 1, the flow will go sonic. It is needed to find M2 at a station 20 ft from station 1. So

L ∗ = 40.81 ft 20 ft, (4.514) 2 − = 20.81 ft (4.515)

fL2∗ 0.0384 (20.81 ft) = 1 , (4.516) D 12 ft = 9.589 (4.517) 1 M 2 1+1.4 (1+1.4) M 2 9.589 = − 2 + ln 2 (4.518) 1.4M 2 2 (1.4) 2+ M 2 (1.4 1) 2 2 − Iterative solution gives

M2 =0.237925 (4.519)

Since energy conservation holds in this flow

u2 u2 h + 2 = h + 1 (4.520) 2 2 1 2 2 2 u2 u1 T2 + = T1 + (4.521) 2cp 2cp 2 ft 2 200 u2 s T2 + = 520 R + 2 (4.522) 2cp  ft 2 6, 015 s2 R 2   u2 T2 + = 523.33 R (4.523) 2cp 2 M2 γRT2 T2 + = 523.33 R (4.524) 2cp

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.6. FLOW WITH HEAT TRANSFER– 117

(γ 1) T 1+ − M 2 = 523.33 R (4.525) 2 2 2   523.33 R T2 = (4.526) (γ−1) 2 1+ 2 M2 523.33 R = − (4.527) (1.4 1) 2 1+ 2 0.237925 = 517.47 R (4.528)

u2 = M2 γRT2, (4.529) p ft2 = 0.237925 1.4 1, 715 (517.47 R) (4.530) s2 R s   ft = 265.2 (4.531) s ρ2u2 = ρ1u1 (4.532) u1 ρ2 = ρ1 , (4.533) u2 lbm 200 ft = 0.2077 s , (4.534) ft3 265.2 ft   s ! lbm = 0.1566 (4.535) ft3

P2 = ρ2RT2, (4.536) lbm ft lbf ft2 = 0.1566 53.34 (517.47 R) (4.537) ft3 lbm R 144 in2       = 30.02 psia (4.538)

T2 P2 s2 s1 = cp ln R ln (4.539) − T1 − P1 ft2 517.47 R ft2 30.02 psia = 6, 015 ln 1, 715 ln (4.540) s2 R 520 R − s2 R 40 psia     ft2 = 462.9 (4.541) s2 R

4.6 Flow with heat transfer–Rayleigh flow

Flow with heat transfer is commonly known as Rayleigh flow. To isolate the effect of heat transfer, the following assumptions will be adopted:

constant area duct • no wall friction • calorically perfect ideal gas •

CC BY-NC-ND. 28 October 2019, J. M. Powers. 118 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

Consequences of heat addition:

stagnation temperature changes • heating drives both subsonic and supersonic flows towards sonic states • cooling drives both subsonic and supersonic flows away from sonic state • heating increases T , decreases P , both subsonic and supersonic • o o cooling decreases T , increases P , both subsonic and supersonic • o o The governing equations are

ρ2u2 = ρ1u1 (4.542) 2 2 ρ2u2 + P2 = ρ1D + P1 (4.543) u2 u2 ρ u A h + 2 = ρ u A h + 1 + q L (4.544) 2 2 2 2 1 1 1 2 wL     γ P P h = o + h (4.545) γ 1 ρ − ρ o −  o  Note that these are a more general case of the equations for a normal shock. One could get equivalents of Rayleigh lines and Hugoniots. The Rayleigh line would be the same as the equations are the same; the Hugoniot would be modified because of the heat transfer term.

If one defines the heat transfer per unit mass of flow q in terms of the wall heat flux qw:

q L q wL (4.546) ≡ ρ1u1A the energy equation becomes

u2 u2 h + 2 = h + 1 + q (4.547) 2 2 1 2 ho2 = ho1 + q (4.548)

q = ho2 ho1 (4.549) q − = To2 To1 (4.550) cp −

With lots of effort very similar to that used for the normal shock equations, expressions can be developed relating the “2” state to the “1” state. If one takes the final “2” state to be sonic 2 and the initial “1” state to be unsubscripted, it is found for the calorically perfect ideal→ ∗ gas that

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.6. FLOW WITH HEAT TRANSFER–RAYLEIGH FLOW 119

2 2 To (γ + 1) M (2+(γ 1) M ) = 2− (4.551) To (1 + γM 2) ∗

Example 4.14 Heat Addition Problem12

ft ◦ Given: Fuel air mixture enters combustion chamber at u1 = 250 s , P1 = 20 psia, T1 = 70 F . The Btu mixture releases 400 lbm

Find: Exit properties u2, P2, T2, heat addition to cause flow to go sonic at exit

Assume: Fuel air mixture behaves just like calorically perfect ideal air

Analysis: Initial state

T1 = 70+460, (4.552) = 530 R (4.553)

c1 = γRT1, (4.554) p ft2 = 1.4 1, 716 (530 R), (4.555) s2 R s   ft = 1, 128.4 (4.556) s u1 M1 = , (4.557) c1 ft 250 s = ft , (4.558) 1, 128.4 s = 0.2216 (4.559)

P1 ρ1 = , (4.560) RT1 lbf lbm ft2 in2 20 in2 32.17 lbfs2 144 ft2 = , (4.561)   ft2   1, 716 s2 R (530 R) lbm  = 0.1019 (4.562) ft3 1 T = T 1+ M 2 , (4.563) o1 1 5 1   1 = (530 R) 1+ 0.22162 , (4.564) 5   = 535.2 R (4.565) 1 3.5 P = P 1+ M 2 , (4.566) o1 1 5 1   12adopted from White, pp. 557-558

CC BY-NC-ND. 28 October 2019, J. M. Powers. 120 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

1 3.5 = (20 psia) 1+ 0.22162 , (4.567) 5   = 20.70 psia (4.568)

Now calculate To∗, the stagnation temperature corresponding to sonic flow

T (γ + 1) M 2 2 + (γ 1) M 2 o1 = 1 − 1 (4.569) T ∗ 2 2 o (1 + γM1 )  T (1.4+1) 0.22162 2+(1.4 1) 0.22162 o1 = − , (4.570) 2 2 To∗ (1 + (1.4)(0 .2216 ))  = 0.2084 (4.571)

To1 T ∗ = , (4.572) o 0.2084 535.2 R = , (4.573) 0.2084 = 2568.3 R (4.574)

Now calculate the effect of heat addition: Btu ft lbf lbm ft q = 400 779 32.17 , (4.575) lbm Btu lbf s2       ft2 = 10.024 106 (4.576) × s2 q To2 = To1 + , (4.577) cp 2 6 ft 10.024 10 2 × s = 535.2+ ft2 , (4.578) 6, 015 s2 R = 2, 201.7 R (4.579) T 2, 201.7 R o2 = , (4.580) To∗ 2, 563.3 R = 0.8573 (4.581) T (γ + 1) M 2 2 + (γ 1) M 2 o2 = 2 − 2 (4.582) T ∗ 2 2 o (1 + γM2 )  (1.4+1) M 2 2+(1.4 1) M 2 0.8573 = 2 − 2 (4.583) 2 2 (1+1 .4M2 )  Computer algebra gives four solutions. For a continuous variation of M, choose the positive subsonic branch. Other branches do have physical meaning.

relevant branch M2 = 0.6380 (4.584) M = 0.6380 (4.585) 2 − M2 = 1.710 (4.586) M = 1.710 (4.587) 2 − Calculate other variables at state 2: − 1 1 T = T 1+ M 2 , (4.588) 2 o2 5 2   CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.6. FLOW WITH HEAT TRANSFER–RAYLEIGH FLOW 121

− 1 1 = (2, 201.7 R) 1+ 0.63802 , (4.589) 5   = 2, 036 R  (4.590)

c2 = γRT2, (4.591) p ft2 = 1.4 1, 716 (2, 036 R), (4.592) s2 R s   ft = 2, 211.6 (4.593) s u2 = M2c2, (4.594) ft = (0.6380) 2, 211.6 , (4.595) s   ft = 1, 411 (4.596) s ρ2u2 = ρ1u1 (4.597) u1 ρ2 = ρ1 , (4.598) u2 lbm 250 ft = 0.1019 s , (4.599) ft3 1, 411 ft   s ! lbm = 0.01806 (4.600) ft3

P2 = ρ2RT2 (4.601) lbm ft2 1 lbf s2 ft2 = 0.01806 1, 716 (2, 036 R) , (4.602) ft3 s2 R 32.17 lbm ft 144 in2       = 13.62 psia (4.603) Is momentum satisfied?

2 2 P2 + ρ2u2 = P1 + ρ1u1 lbf 144 in2 lbm ft 2 1 lbf s2 13.62 + 0.01806 1, 411 in2 ft2 ft3 s 32.17 lbm ft           lbf 144 in2 lbm ft 2 1 lbf s2 = 20 + 0.1019 250 in2 ft2 ft3 s 32.17 lbm ft           lbf lbf 3, 078.97 = 3077.97 close! ft2 ft2 Entropy Change

T2 P2 s2 s1 = cp ln R ln (4.604) − T1 − P1 ft2 2, 036 R ft2 13.62 psia = 6, 015 ln 1, 716 ln (4.605) s2 R 530 R − s2 R 20 psia         = 8, 095.38 ( 659.28) , (4.606) − − ft2 = 8, 754.66 (4.607) s2 R ft2 1 Btu 1 lbf s2 = 8, 754.66 (4.608) s2 R 779 ft lbf 32.17 lbm ft       CC BY-NC-ND. 28 October 2019, J. M. Powers. 122 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

Btu = 0.3493 (4.609) lbm R Second Law q s s (4.610) 2 − 1 ≥ T Btu 400 Btu 0.3493 lbm (4.611) lbm R ≥ 2, 036 R Btu Btu 0.3493 0.1965 yes! (4.612) lbm R ≥ lbm R maximum heat release

q = c (T ∗ T ) (4.613) max p o − o1 ft2 1 Btu 1 lbf s2 = 6, 015 (2, 568.3 R 535.2 R) (4.614) s2 R − 779 ft lbf 32.17 lbm ft       Btu q = 488 (4.615) max lbm

4.7 Numerical solution of the shock tube problem

A detailed development is given in lecture for the numerical solution to the Riemann or shock tube problem. The equations are first posed in the general conservative form: ∂q ∂ + (f(q))=0. (4.616) ∂t ∂x Here q and f vector functions of length N = 3; further f is itself a function of q. The equations are discretized so that

q(x, t) qn, (4.617) → i f(q(x, t)) f(qn). (4.618) → i 4.7.1 One-step techniques A brief discussion of finite difference techniques is given in lecture. The most tempting technique is a first order forward difference in time, central difference in space technique which yields the finite difference relation:

n+1 n ∆t n n qi = qi f(qi+1) f(qi 1) . (4.619) − 2∆x − − Unfortunately this method is unstable. 

CC BY-NC-ND. 28 October 2019, J. M. Powers. 4.7. NUMERICAL SOLUTION OF THE SHOCK TUBE PROBLEM 123

4.7.2 Lax-Friedrichs technique A robustly stable first order method is found int the Lax-Friedrichs method.

n+1 1 n n ∆t n n qi = qi 1 + qi+1 f(qi+1) f(qi 1) . (4.620) 2 − − 2∆x − −   4.7.3 Lax-Wendroff technique The two-step Lax-Wendroff discretization is as follows

at a given time step estimate q at the i +1/2 cell interface: • 1 qn = qn + qn , (4.621) i+1/2 2 i i+1  n+1/2 use central differencing (about i +1/2) to step forward ∆t/2 so that qi+1/2 can be • estimated:

∆t/2 qn+1/2 = qn f(qn ) f(qn) . (4.622) i+1/2 i+1/2 − ∆x i+1 − i  use central differencing (about i) to step forward ∆t, evaluating f at the i 1/2 and • ± n +1/2 steps:

n+1 n ∆t n+1/2 n+1/2 qi = qi f(qi+1/2 ) f(qi 1/2 ) . (4.623) − ∆x − −  

CC BY-NC-ND. 28 October 2019, J. M. Powers. 124 CHAPTER 4. ONE-DIMENSIONAL COMPRESSIBLE FLOW

CC BY-NC-ND. 28 October 2019, J. M. Powers. Chapter 5

Steady supersonic two-dimensional flow

Suggested Reading:

Liepmann and Roshko, Chapter 4: pp. 84-123 Hughes and Brighton, Chapter 8: pp. 208-230 Shapiro, Chapters 9-16: pp. 265-609 White, Chapter 9: pp. 559-581

This chapter will discuss two-dimensional flow of a compressible fluid. The following topics will be covered:

presentation of isentropic two-dimensional flow equations • oblique shocks • Prandtl-Meyer rarefactions • flow over an airfoil •

Assume for this chapter:

∂ 0; steady flow • ∂t ≡ w 0, ∂ 0; two-dimensional flow • ≡ ∂z ≡ no viscous stress or heat conduction, so isentropic except through shocks • calorically perfect ideal gas • 125 126 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

5.1 Two-dimensional equations

With the assumptions of above the following equations govern the flow away from shock discontinuities:

5.1.1 Conservative form

∂ ∂ (ρu)+ (ρv) = 0 (5.1) ∂x ∂y ∂ ∂ ρu2 + P + (ρuv) = 0 (5.2) ∂x ∂y ∂ ∂ (ρvu)+ ρv2 + P = 0 (5.3) ∂x ∂y ∂ 1 P ∂ 1 P  ρu e + u2 + v2 + + ρv e + u2 + v2 + = 0 (5.4) ∂x 2 ρ ∂y 2 ρ         1 P e = + e (5.5) γ 1 ρ o − 5.1.2 Non-conservative form

∂ρ ∂ρ ∂u ∂v u + v + ρ + = 0 (5.6) ∂x ∂y ∂x ∂y     ∂u ∂u ∂P ρ u + v + = 0 (5.7) ∂x ∂y ∂x   ∂v ∂v ∂P ρ u + v + = 0 (5.8) ∂x ∂y ∂y   ∂e ∂e ∂u ∂v ρ u + v + P + = 0 (5.9) ∂x ∂y ∂x ∂y     1 P e = + e (5.10) γ 1 ρ o − 5.2 Mach waves

Mach waves are small acoustic disturbances in a flow field. Recall that small disturbances propagate at the ambient sound speed. Let’s consider a small sphere moving at u1 through a fluid with ambient sound speed co.

u

CC BY-NC-ND. 28 October 2019, J. M. Powers. 5.3. OBLIQUE SHOCK WAVES 127

u >c , supersonic flow, larger region still unaware of sphere • 1 o Consider that in time ∆t, the sphere will move u1∆t and the wave will propagate will be felt by a circle with radius co∆t, see Figure 5.1

co Δt co Δt

u1 < co u1 = co

subsonic sonic flow flow

u1 Δ t

u1 Δ t β = arcsin (1/M1) zone of silence

co Δt

u1 > co

supersonic flow

u1 Δ t

Figure 5.1: Acoustic disturbance sketch

From the geometry,

c ∆t c 1 sin β = o = o = (5.11) u1∆t u1 M1 1 β = arcsin (5.12) M  1  5.3 Oblique shock waves

An oblique shock is a shock which is not normal to the incoming flow field. It can be shown that in the limiting case as the oblique shock strength goes to zero, the oblique shock wave becomes a Mach wave, as described in the previous section.

Oblique waves can be understood by considering the following problem.

CC BY-NC-ND. 28 October 2019, J. M. Powers. 128 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

Given: a straight wedge inclined at angle θ to the horizontal • a freestream flow parallel to the horizontal with known velocity v = u i +0 j • 1 known freestream pressure and density of P and ρ • 1 1 steady flow of a calorically perfect ideal gas (this can be relaxed and one can still find • oblique shocks) Find: angle of shock inclination β • downstream pressure and density P , ρ • 2 2 Similar to the piston problem, the oblique shock problem is easiest analyzed if we instead consider β as known • θ as unknown • They are best modeled in a two-dimensional coordinate system with axes parallel and perpendicular to the shock, see Figure 5.2, so that x =x ˜ sin β +˜y cos β (5.13) y = x˜ cos β +˜y sin β (5.14) − u =u ˜ sin β +˜v cos β (5.15) v = u˜ cos β +˜v sin β (5.16) − Consequently, in this coordinate system, the freestream is two-dimensional. It is easily shown that the equations of motion are invariant under a rotation of axes, so that

∂ ∂ (ρu˜)+ (ρv˜) = 0 (5.17) ∂x˜ ∂y˜ ∂ ∂ ρu˜2 + P + (ρu˜v˜) = 0 (5.18) ∂x˜ ∂y˜ ∂ ∂ (ρv˜u˜)+ ρv˜2 + P = 0 (5.19) ∂x˜ ∂y˜ ∂ 1 P ∂ 1 P  ρu˜ e + u˜2 +˜v2 + + ρv˜ e + u˜2 +˜v2 + = 0 (5.20) ∂x˜ 2 ρ ∂y˜ 2 ρ         1 P e = + e (5.21) γ 1 ρ o − To analyze oblique shocks, we make one additional assumption

CC BY-NC-ND. 28 October 2019, J. M. Powers. 5.3. OBLIQUE SHOCK WAVES 129

oblique shock wave

tan β = u∼ / v∼ ∼ ∼ 1 1 tan (β−θ) = u2 / v2 u~ ~ ~v 2 ~ u1 2 ~ v1 v2 P1 β−θ β P2 ρ 1 u 1 ~u ρ2 2 unshocked shocked freestream flow (supersonic) ~ y β y θ wedge x

~ x

Figure 5.2: Oblique Shock Schematic

∂ =0 • ∂y˜

Note however that, contrary to one-dimensional flow we will not enforcev ˜ = 0, so

v˜ =0 • 6 ∂ d Consequently, all variables are a function ofx ˜ at most and ∂x˜ = dx˜ . The governing equations reduce to

d (ρu˜) = 0 (5.22) dx˜ d ρu˜2 + P = 0 (5.23) dx˜ d (ρv˜u˜) = 0 (5.24) dx˜ d 1 P ρu˜ e + u˜2 +˜v2 + = 0 (5.25) dx˜ 2 ρ     1 P e = + e (5.26) γ 1 ρ o − Integrate and apply freestream conditions

CC BY-NC-ND. 28 October 2019, J. M. Powers. 130 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

ρ2u˜2 = ρ1u˜1 (5.27) 2 2 ρ2u˜2 + P2 = ρ1u˜1 + P1 (5.28)

ρ2v˜2u˜2 = ρ1v˜1u˜1 (5.29) 1 P 1 P ρ u˜ e + u˜2 +˜v2 + 2 = ρ u˜ e + u˜2 +˜v2 + 1 (5.30) 2 2 2 2 2 2 ρ 1 1 1 2 1 1 ρ  2   1   1  P e = + e (5.31) γ 1 ρ o − Now using the mass equation, they ˜ momentum equation reduces to

v˜2 =v ˜1 (5.32)

Using this result and the mass a state equations gives

ρ2u˜2 = ρ1u˜1 (5.33) 2 2 ρ2u˜2 + P2 = ρ1u˜1 + P1 (5.34) 1 P 1 P 1 P 1 P 2 + u˜2 + 2 = 1 + u˜2 + 1 (5.35) γ 1 ρ 2 2 ρ γ 1 ρ 2 1 ρ − 2 2 − 1 1 These are exactly the equations which describe a normal shock jump. All our old results apply in this coordinate system with the additional stipulation that the component of velocity tangent to the shock is constant.

Recall our solution for one-dimensional shocks in a calorically perfect ideal gas:

1 1 γ 1 2γ P = − 1+ 1 (5.36) ρ ρ γ +1 (γ 1) D2 ρ 2 1  − 1 

For this problem D =u ˜1 so

1 1 γ 1 2γ P = − 1+ 1 (5.37) ρ ρ γ +1 (γ 1)u ˜2 ρ 2 1  − 1 1  With the freestream Mach number normal to the wave defined as

u˜2 M 2 1 (5.38) 1n ≡ γ P1 ρ1

CC BY-NC-ND. 28 October 2019, J. M. Powers. 5.3. OBLIQUE SHOCK WAVES 131 we get

ρ γ 1 2 1 = − 1+ (5.39) ρ γ +1 (γ 1) M 2 2  − 1n  and since from mass ρ1 = u˜2 ρ2 u˜1

u˜ γ 1 2 2 = − 1+ (5.40) u˜ γ +1 (γ 1) M 2 1  − 1n  Now for our geometry

u˜ tan β = 1 (5.41) v˜1 u˜ u˜ tan(β θ)= 2 = 2 (5.42) − v˜2 v˜1 u˜ tan(β θ) so 2 = − (5.43) u˜1 tan β thus tan(β θ) γ 1 2 − = − 1+ (5.44) tan β γ +1 (γ 1) M 2  − 1n  Now note that

2 2 2 M1n = M1 sin β (5.45) so tan(β θ) γ 1 2 − = − 1+ (5.46) tan β γ +1 (γ 1) M 2 sin2 β  − 1  tan(β θ) γ 1 (γ 1) M 2 sin2 β +2 − = − − 1 (5.47) tan β γ +1 (γ 1) M 2 sin2 β  1  − 2 2 (γ 1) M1 sin β +2 tan(β θ) = tan β − 2 2 (5.48) − (γ + 1) M1 sin β 2 2 tan β tan θ (γ 1) M1 sin β +2 − = tan β − 2 2 χ (5.49) 1+tan θ tan β (γ + 1) M1 sin β ≡ tan β tan θ = χ + χ tan θ tan β (5.50) − tan β χ = tan θ (1 + χ tan β) (5.51) − tan β χ tan θ = − (5.52) 1+ χ tan β

CC BY-NC-ND. 28 October 2019, J. M. Powers. 132 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

With a little more algebra and trigonometry this reduces to

2 2 M1 sin β 1 tan θ = 2cotβ 2 − (5.53) M1 (γ + cos2β)+2

Given M1, γ and β, this equation can be solved to find θ the wedge angle. It can be inverted to form an equation cubic in sin β to solve explicitly for β. Figure 5.3 gives a plot of oblique shock angle β versus wedge angle θ.

β expanded view of β−θ plane

strong branch β 90 (post-shock subsonic) 75

Μ1 = 2 2nd law 50 violation 60 Μ = 3 maximum wedge 25 1 angle for attached Μ = ∞ oblique shock 1 θ -40 -20 20 40

Μ1 = 2 30 weak branch -25 (post shock supersonic, primarily) Μ = 3 1 2nd Law -50 Violation Μ = ∞ 1 γ = 7/5 -75 0 θ 0 10 20 30 40 50

Figure 5.3: Shock angle β versus wedge angle θ

Note the following features:

for a given θ < θ , there exist two β’s • max – lower β is weak solution 1 limθ 0 β = arcsin , a Mach wave ∗ → M relevant branch for most external flows, matches in far-field to acoustic wave, ∗ can exist in internal flows 2 u˜2+˜v2 total Mach number primarily supersonic, M2 = 2 > 1 for nearly all 0 < ∗ c2 θ < θmax 2 u˜2 normal Mach number subsonic, M2n = 2 < 1 ∗ c2

CC BY-NC-ND. 28 October 2019, J. M. Powers. 5.3. OBLIQUE SHOCK WAVES 133

– higher β is the strong solution π limθ 0 β = , a normal shock wave ∗ → 2 relevant branch for some internal flows ∗ 2 u˜2+˜v2 total Mach number completely subsonic, M2 = 2 < 1 for all 0 <θ<θmax ∗ c2 2 u˜2 normal Mach number subsonic, M2n = 2 < 1 ∗ c2 for θ > θ , no solution exists; shock becomes detached • max Consider fixed θ, increasing freestream Mach number M , see Figure 5.4 • 1

– 0 < M1 < 1, subsonic incoming flow, no shocks continuous pressure variation

– 1 < M1 < M1a, supersonic incoming flow, detached curved oblique shock – M < M < , supersonic incoming flow, attached straight oblique shock 1a 1 ∞ – as M1 , β β →∞ → ∞ Consider fixed supersonic freestream Mach number M , increasing θ, see Figure 5.5 • 1 – θ 0, Mach wave, negligible disturbance ∼ – small θ, small β, small pressure, density rise – medium θ, medium β, moderate pressure and density rise – large θ, curved detached shock, large pressure and density rise

Example 5.1 Oblique Shock Example

◦ Given: Air flowing over a wedge, θ = 20 , P1 = 100 kPa, T1 = 300 K, M1 =3.0

Find: Shock angle β and downstream pressure and temperature P2, T2.

Assume: calorically perfect ideal gas

Analysis: First some preliminaries:

J m c = γRT = (1.4) 287 (300 K) = 347.2 (5.54) 1 1 kg K s s   p m m u = M c = (3.0) 347.2 =1, 041.6 (5.55) 1 1 1 s s P1 100, 000 Pa  kg ρ1 = = =1.1614 3 (5.56) RT1 J m 287 kg K (300 K)   CC BY-NC-ND. 28 October 2019, J. M. Powers. 134 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

continuous pressure shock approaching detached variation wedge from infinity oblique shock

M1 < 1 M1 ~ 1 1 < M1 < M 1a subsonic, shockless flow slightly supersonic flow moderately supersonic flow

attached attached oblique shock oblique shock at limiting wave angle

βmin

M1 ->∞ M1a < M1 < ∞

Figure 5.4: Shock wave patterns as incoming Mach number varied

Now find the wave angle:

2 2 M1 sin β 1 tan θ = 2cotβ 2 − (5.57) M1 (γ + cos2β)+2 2 2 ◦ 3.0 sin β 1 tan20 = 2cotβ − (5.58) 3.02 (1.4+cos2β)+2 (5.59)

Three solutions:

weak oblique shock; common β = 37.76◦ (5.60) strong oblique shock; rare β = 82.15◦ (5.61) second law violating “rarefaction” shock β = 9.91◦ (5.62) −

1. Weak Oblique Shock

m ◦ m u˜ = u sin β = 1, 041.6 sin 37.76 = 637.83 (5.63) 1 1 s s  m  ◦ m v˜ = u cos β = 1, 041.6 cos37.76 = 823.47 (5.64) 1 1 s s  u˜  637.83 m M = 1 = s =1.837 (5.65) 1n c 347.2 m  1   s  ρ γ 1 2 1 = − 1+ (5.66) ρ γ +1 (γ 1) M 2 2  − 1n  CC BY-NC-ND. 28 October 2019, J. M. Powers. 5.3. OBLIQUE SHOCK WAVES 135

M1 !xed, supersonic, P 1, ρ1 !xed

M1 M1 M1 M1

Mach wave attached shocks detached shock

small disturbance moderate disturbances large disturbance

Figure 5.5: Shock wave patterns as wedge angle varied

kg 1.1614 3 1.4 1 2 m = − 1+ =0.413594 (5.67) ρ 1.4+1 (1.4 1)1.8372 2  −  kg 1.1614 3 kg ρ = m =2.8081 (5.68) 2 0.41359 m3 ρ2u˜2 = ρ1u˜1 (5.69) kg m ρ u˜ 1.1614 m3 637.83 s m u˜ = 1 1 = = 263.80 (5.70) 2 ρ   kg s 2 2.8081 m3  m v˜ =v ˜ = 823.47 (5.71) 2 1 s u2 =u ˜2 sin β +˜v2 cos β (5.72) v = u˜ cos β +˜v sin β (5.73) 2 − 2 2 m ◦ m ◦ m u = 263.80 sin 37.76 + 823.47 cos37.76 = 812.56 (5.74) 2 s s s  m ◦  m ◦ m v = 263.80 cos37.76 + 823.47 sin 37.76 = 295.70 (5.75) 2 − s s s     v check on wedge angle θ = arctan 2 (5.76) u  2  295.70 m = arctan s = 19.997◦ (5.77) 812.56 m  s  P = P + ρ u˜2 ρ u˜2 (5.78) 2 1 1 1 − 2 2 kg m 2 kg m 2 P = 100, 000 Pa + 1.1614 637.83 2.8081 263.80 (5.79) 2 m3 s − m3 s         P2 = 377, 072 Pa (5.80) P2 377, 072 Pa T2 = = = 467.88 K (5.81) ρ2R kg J 2.8081 m3 287 kg K    J m c = γRT = (1.4) 287 (467.88 K) = 433.58 (5.82) 2 2 kg K s s   p CC BY-NC-ND. 28 October 2019, J. M. Powers. 136 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

m u˜2 263.8 s M2n = = m =0.608 (5.83) c2 433.58 s m 2 m 2 u2 + v2 812.56 + 295.7 M = 2 2 = s s =1.994 (5.84) 2 c q 433.58 m p 2  s  T2 P2 s2 s1 = cp ln R ln (5.85) − T1 − P1 J 467.88 K J 377, 072 Pa = 1, 004.5 ln 287 ln (5.86) kg K 300 K − kg K 100, 000 Pa     J s s = 65.50 (5.87) 2 − 1 kg K

2. Strong Oblique Shock

m m u˜ = u sin β = 1, 041.6 sin 82.15◦ =1, 031.84 (5.88) 1 1 s s  m ◦ m v˜ = u cos β = 1, 041.6 cos82.15 = 142.26 (5.89) 1 1 s s  u˜  1, 031.84 m M = 1 = s =2.972 (5.90) 1n c 347.2 m  1   s  ρ γ 1 2 1 = − 1+ (5.91) ρ γ +1 (γ 1) M 2 2  − 1n  kg 1.1614 3 1.4 1 2 m = − 1+ =0.26102 (5.92) ρ 1.4+1 (1.4 1)2.9722 2  −  kg 1.1614 3 kg ρ = m =4.4495 (5.93) 2 0.26102 m3 ρ2u˜2 = ρ1u˜1 (5.94) kg m ρ u˜ 1.1614 m3 1, 031.84 s m u˜ = 1 1 = = 269.33 (5.95) 2 ρ   kg s 2 4.4495 m3  m v˜ =v ˜ = 142.26 (5.96) 2 1 s u2 =u ˜2 sin β +˜v2 cos β (5.97) v = u˜ cos β +˜v sin β (5.98) 2 − 2 2 m ◦ m ◦ m u = 269.33 sin 82.15 + 142.26 cos82.15 = 286.24 (5.99) 2 s s s  m ◦  m ◦ m v = 269.33 cos82.15 + 142.26 sin 82.15 = 104.14 (5.100) 2 − s s s     v check on wedge angle θ = arctan 2 (5.101) u  2  104.14 m = arctan s = 19.99◦ (5.102) 286.24 m  s  P = P + ρ u˜2 ρ u˜2 (5.103) 2 1 1 1 − 2 2 kg m 2 kg m 2 P = 100, 000 Pa + 1.1614 1, 031.84 4.4495 269.33 (5.104) 2 m3 s − m3 s         P2 =1, 013, 775 Pa (5.105)

CC BY-NC-ND. 28 October 2019, J. M. Powers. 5.3. OBLIQUE SHOCK WAVES 137

P2 1, 013, 775 Pa T2 = = = 793.86 K (5.106) ρ2R kg J 4.4495 m3 287 kg K    J m c = γRT = (1.4) 287 (793.86 K) = 564.78 (5.107) 2 2 kg K s s   p m u˜2 269.33 s M2n = = m =0.477 (5.108) c2 564.78 s m 2 m 2 u2 + v2 286.24 + 104.14 M = 2 2 = s s =0.539 (5.109) 2 c q 564.78 m p 2  s  T2 P2 s2 s1 = cp ln R ln (5.110) − T1 − P1 J 793.86 K J 1, 013, 775 Pa = 1, 004.5 ln 287 ln (5.111) kg K 300 K − kg K 100, 000 Pa     J s s = 312.86 (5.112) 2 − 1 kg K 3. “Rarefaction Shock”

m m u˜ = u sin β = 1, 041.6 sin ( 9.91◦)= 179.26 (5.113) 1 1 s − − s  m  ◦ m v˜ = u cos β = 1, 041.6 cos( 9.91 )=1, 026.06 (5.114) 1 1 s − s  u˜  179.26 m M = 1 = − s = 0.5163 (5.115) 1n c 347.2 m −  1   s  ρ γ 1 2 1 = − 1+ (5.116) ρ γ +1 (γ 1) M 2 2  − 1n  kg 1.1614 m3 1.4 1 2 = − 1+ 2 =3.2928 (5.117) ρ2 1.4+1 (1.4 1) ( 0.5163) ! − − kg 1.1614 3 kg ρ = m =0.3527 (5.118) 2 3.2928 m3 ρ2u˜2 = ρ1u˜1 (5.119) kg m ρ u˜ 1.1614 m3 179.26 s m u˜ = 1 1 = − = 590.27 (5.120) 2 ρ   kg − s 2 0.3527 m3  m v˜ =v ˜ =1, 026.06 (5.121) 2 1 s u2 =u ˜2 sin β +˜v2 cos β (5.122)

v2 = u˜2 cos β +˜v2 sin β (5.123) m m − m u = 590.27 sin ( 9.91◦)+ 1, 026.06 cos( 9.91◦)=1, 112.34 (5.124) 2 − s − s − s  m ◦  m ◦ m v = 590.27 cos( 9.91 )+ 1, 026.06 sin ( 9.91 ) = 404.88 (5.125) 2 − − s − s − s     v check on wedge angle θ = arctan 2 (5.126) u2 m   404.88 ◦ = arctan s = 20.00 (5.127) 1, 112.34 m  s 

CC BY-NC-ND. 28 October 2019, J. M. Powers. 138 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

P = P + ρ u˜2 ρ u˜2 (5.128) 2 1 1 1 − 2 2 kg m 2 kg m 2 P = 100, 000 Pa + 1.1614 179.26 0.3527 590.27 (5.129) 2 m3 − s − m3 − s         P2 = 14, 433 Pa (5.130) P2 14, 433 Pa T2 = = = 142.59 K (5.131) ρ2R kg J 0.3527 m3 287 kg K    J m c = γRT = (1.4) 287 (142.59 K) = 239.36 (5.132) 2 2 kg K s s   p m u˜2 590.27 s M2n = = − m = 2.47 (5.133) c2 239.36 s − m 2 m 2 u2 + v2 1, 112.34 + 404.88 M = 2 2 = s s =4.95 (5.134) 2 c q 239.36 m p 2  s  T2 P2 s2 s1 = cp ln R ln (5.135) − T1 − P1 J 142.59 K J 14, 433 Pa = 1, 004.5 ln 287 ln (5.136) kg K 300 K − kg K 100, 000 Pa     J s s = 191.5 (5.137) 2 − 1 − kg K

5.4 Small disturbance theory

By taking a Taylor series expansion of the relationship between β and θ about θ = 0, for fixed M1 and γ it can be shown that

1 γ +1 M 4 tan β = + 1 θ θ<< 1 (5.138) M 2 1 4 (M 2 1)2 1 − 1 − Note that when θ =p 0 that 1 tan β = (5.139) M 2 1 1 − 1 tan2 β =p (5.140) M 2 1 1 − sin2 β sin2 β 1 2 = 2 = 2 (5.141) cos β 1 sin β M1 1 − 1− 2 2 sin β M1 2 = 1 (5.142) 1 sin β 1 2 − − M1

CC BY-NC-ND. 28 October 2019, J. M. Powers. 5.4. SMALL DISTURBANCE THEORY 139

1 sin β = (5.143) M1 1 β = arcsin (5.144) M  1  After a good deal of algebra and trigonometry, it can also be shown that the pressure change, change in velocity magnitude, w, and change in entropy for flow over a thin wedge is

P P γM 2 2 − 1 = 1 θ (5.145) 2 P1 M 1 1 − w w θ 2 − 1 = p (5.146) 2 w1 − M1 1 s s − 2 p− 1 θ3 (5.147) s1 ∼ In terms of changes,

∆P γM 2 = 1 ∆θ (5.148) 2 P1 M 1 1 − ∆w ∆θ =p (5.149) 2 w1 − M 1 1 − ∆s p ∆θ3 (5.150) s1 ∼ Note that a small positive ∆θ gives rise to an increase in pressure • a decrease in velocity magnitude • a very small change in entropy • Figure 5.6 shows the pattern of waves that one obtains when subjecting a flow to a series of small turns and the pattern that evolves as the turning radius is shrunk.

compression waves converge • expansion waves diverge • convergence of compression waves leads to region of rapid entropy rise–shock formation • divergence of pressure waves leads to no shock formation in expansion • This has an analog in one-dimensional unsteady flow. Consider a piston with an initial velocity of zero accelerating into a tube

CC BY-NC-ND. 28 October 2019, J. M. Powers. 140 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

Wave convergence; Shock Formation; Breakdown of Isentropic Assumption

Isentropic Compression

Isentropic Expansion

Wave divergence;

No shock formation

Oblique Shock

Prandtl-Meyer Expansion

Figure 5.6: Wave pattern and streamlines for flows undergoing a series of small turns and for sudden turns

lead compression wave travels at sound speed • lead wave increases temperature (and sound speed) of disturbed flow • each successive acoustic wave travels faster than lead wave • eventually acoustic waves catch and form a shock • Consider a piston with zero initial velocity which decelerates

lead expansion wave travels at sound speed • lead wave decreases temperature (and sound speed) of disturbed flow • each successive acoustic wave travels slow than lead wave • no shock formation • A schematic for these one-dimensional unsteady flows is shown in Figure 5.7

CC BY-NC-ND. 28 October 2019, J. M. Powers. 5.5. CENTERED PRANDTL-MEYER RAREFACTION 141

t t accelerating piston shock path

decelerating slow, acoustic piston lead wave path

x x

P , ρ 1 1 P1, ρ1

t suddenly t accelerated piston path

shock locus

suddenly decelerated piston path Prandtl-Meyer expansion fan

x x

Figure 5.7: Schematic of compression and expansion waves for one-dimensional unsteady piston-driven flow

5.5 Centered Prandtl-Meyer rarefaction

If we let ∆θ 0, the entropy changes become negligibly small relative to pressure and → velocity changes, and the flow is isentropic. The relations can be replaced by differential relations:

dP γM 2 = dθ (5.151) P √M 2 1 dw −dθ = (5.152) w −√M 2 1 ds − 0 (5.153) s ∼

CC BY-NC-ND. 28 October 2019, J. M. Powers. 142 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

Recall now that for adiabatic flow T γ 1 o =1+ − M 2 (5.154) T 2 γRT γ 1 o =1+ − M 2 (5.155) γRT 2 c2 γ 1 o =1+ − M 2 (5.156) c2 2 1 γ 1 − 2 c = c 1+ − M 2 (5.157) o 2  3  c γ 1 − 2 dc = o 1+ − M 2 (γ 1) MdM (5.158) − 2 2 −   γ 1 dc − MdM = 2 (5.159) c γ 1 2 − 1+ −2 M Also

w = cM (5.160) dw = cdM + Mdc (5.161) dw dM dc = + (5.162) w M c γ 1 dw dM − MdM = 2 (5.163) γ 1 2 w M − 1+ −2 M dw 1 dM = (5.164) γ 1 2 w 1+ −2 M M dθ 1 dM = γ 1 (5.165) − √M 2 1 1+ − M 2 M − 2 √M 2 1 dM dθ = − (5.166) M γ 1 2 − 1+ −2 M Now positive θ corresponds to compression and negative θ corresponds to expansion. Let’s define ν so positive ν gives and expansion.

ν θ + θ (5.167) ≡ − o dν = dθ (5.168) − Now integrate the expression

√M 2 1 dM dν = − (5.169) γ 1 2 M 1+ −2 M

CC BY-NC-ND. 28 October 2019, J. M. Powers. 5.5. CENTERED PRANDTL-MEYER RAREFACTION 143

Let ν = 0 correspond to M = 1. This effectively selects θo

γ +1 1 γ 1 2 1 2 ν(M)= tan− − (M 1) tan− √M 1 (5.170) γ 1 γ +1 − − − r − r The function ν(M) is called the Prandtl-Meyer function. It is plotted in Figure 5.8. Many texts tabulate the Prandtl-Meyer function. For a known turning angle, one can find the Mach number. As the flow is entirely isentropic, all other flow variables can be obtained through the isentropic relations. Note:

π γ+1 As M , ν 2 γ 1 1 , corresponds to vacuum conditions • →∞ → − − q  given ν, one can calculate M • isentropic relations give P , ρ, T , etc. • Prandtl-Meyer function tabulated in many texts •

ν (ο) ο νmax = 130.5

120

100

80

60 γ = 7/5

10 20 30 40 M

Figure 5.8: Prandtl-Meyer function

Example 5.2 Centered Expansion

m Given: Calorically perfect, ideal air with P1 = 100 kPa, T1 = 300 K, u1 = 500 s , turned through a 30◦ expansion corner.

Find: Fluid properties after the expansion

Analysis:

CC BY-NC-ND. 28 October 2019, J. M. Powers. 144 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

P1 100 kPa kg ρ1 = = =1.1614 3 (5.171) RT1 J m 0.287 kg K (300 K)   J m c = γRT = 1.4 287 (300 K) = 347.2 (5.172) 1 1 kg K s s   p m u1 500 s M1 = = m =1.4401 (5.173) c1 347.2 s γ 1 T = T 1+ − M 2 (5.174) o 1 2 1   1 T = 300 K 1+ 1.44012 = 424.43 K (5.175) o 5   γ γ 1 γ−1 P = P 1+ − M 2 (5.176) o 1 2 1   1 3.5 P = 100 kPa 1+ 1.44012 = 336.828 kPa (5.177) o 5   Now calculate the Prandtl-Meyer function for the freestream:

γ +1 − γ 1 − ν(M )= tan 1 − (M 2 1) tan 1 M 2 1 (5.178) 1 γ 1 γ +1 1 − − 1 − r − r q 1.4+1 − 1.4 1 − ν(M )= tan 1 − (1.44012 1) tan 1 1.44012 1 (5.179) 1 1.4 1 1.4+1 − − − r − r ν(M1)=0p.177138 rad (5.180) ◦ 180 ◦ ν(M )=0.177138 rad = 10.1493 (5.181) 1 π rad The interpretation here is that an initially sonic flow would have had to had turned 10.1493◦ to achieve a Mach number of M1 =1.4401.

Now add on the actual turning:

◦ ν(M2)= ν(M1)+30 (5.182) ◦ ◦ ◦ ν(M2)=10.1493 + 30 = 40.1493 (5.183) ◦ π rad ν(M )=40.1493 =0.700737 rad (5.184) 2 180◦

A trial and error solution gives the M2 which corresponds to ν(M2)=0.700737 rad:

1.4+1 − 1.4 1 − 0.700737 rad = tan 1 − (M 2 1) tan 1 M 2 1 (5.185) 1.4 1 1.4+1 2 − − 2 − r − r q M2 =2.54431 (5.186) − γ 1 1 T = T 1+ − M 2 (5.187) 2 o 2 2  −  1 1 T = 424.43 K 1+ 2.543312 = 189.4 K (5.188) 2 5   CC BY-NC-ND. 28 October 2019, J. M. Powers. 5.6. WAVE INTERACTIONS AND REFLECTIONS 145

− γ γ 1 γ−1 P = P 1+ − M 2 (5.189) 2 o 2 2  −  1 3.5 P = 336.828 kPa 1+ 2.543312 = 18.43 kPa (5.190) 2 5   P2 18.43 kPa kg ρ2 = = =0.3390 3 (5.191) RT2 J m 0.287 kg K (189.4 K)   J m c = γRT = 1.4 287 (189.4 K) = 275.87 (5.192) 2 2 kg K s s   p m m w = M c =2.54431 275.87 = 701.89 (5.193) 2 2 2 s s  

5.6 Wave interactions and reflections

Shocks and rarefactions can intersect and reflect in a variety of ways.

5.6.1 Oblique shock reflected from a wall An oblique shock which reflects from a wall is represented in Figure 5.9.

Μ 1 Μ2 Μ3

β1 β2

P3 P 2 interior streamline P1 pressure field

P3

P1 wall pressure

Figure 5.9: Reflection of an oblique shock from a wall

Note:

CC BY-NC-ND. 28 October 2019, J. M. Powers. 146 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

analysis just that of two oblique shocks • flow always turns to be parallel to wall • angle of incidence not equal angle of reflection due to non-linear effects • interior pressure profile has two steps • wall pressure profile has single step • P > 2P , that is the pressure is higher than that obtained in the acoustic limit • 2 1 5.6.2 Oblique shock intersection Two oblique shocks intersect as sketched in Figure 5.10

Μ2

Μ1 Μ3

Μ2

Figure 5.10: Interaction of two oblique shocks

Note:

flow always turns to be parallel to wall • when shocks intersect, flow turns again to be parallel to itself • 5.6.3 Shock strengthening A flow turned by a corner through an oblique shock can be strengthened by a second turn as sketched in Figure 5.11 Note: three new waves generated

CC BY-NC-ND. 28 October 2019, J. M. Powers. 5.7. SUPERSONIC FLOW OVER AIRFOILS 147

slipstream

rarefaction

Figure 5.11: Shock Strengthening Sketch

strengthened shock • slipstream in which pressures match, velocity directions match, but velocity magnitudes • differ

weak rarefaction wave • 5.6.4 Shock weakening A flow turned by a corner through an oblique shock can be weakened by a second turn as sketched in Figure 5.12

Figure 5.12: Shock Weakening Sketch

5.7 Supersonic flow over airfoils

The standard problem in flow over an airfoil is to determine the lift and the drag. While in actual design it is the magnitude of the lift force FL, and drag force FD, that is most crucial,

CC BY-NC-ND. 28 October 2019, J. M. Powers. 148 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

there exists dimensionless numbers the lift coefficient CL and the drag coefficient CD which give good relative measures of airfoil performance.

FL CL 1 2 (5.194) ≡ 2 ρ1u1A FD CD 1 2 (5.195) ≡ 2 ρ1u1A Though this is the traditional formula, it is probably not the best for interpreting how the forces vary when flight speed is varied. This is because when u1, flight speed is varied both numerator and denominator change. To remedy this, we can instead scale by the ambient sound speed to define a dimensionless lift force F and dimensionless drag force F : ∗L ∗D

FL F L 2 (5.196) ∗ ≡ ρ1c1A FD F D 2 (5.197) ∗ ≡ ρ1c1A

5.7.1 Flat plate at angle of attack

The simplest problem is that of a flat plate at angle of attack αo. A schematic is illustrated in Figure 5.13. Note:

far field Mach P1 leading edge wave T1 rarefaction M1

P2

αο slipstream P ' leading edge2 shock

far field Mach wave

Figure 5.13: Supersonic Flow over a Flat Plate

flow over the top is turned through an isentropic rarefaction to P • 2

CC BY-NC-ND. 28 October 2019, J. M. Powers. 5.7. SUPERSONIC FLOW OVER AIRFOILS 149

flow over the bottom is turned through an oblique shock to P ′ • 2

Since P ′ >P , there is both lift and drag forces! • 2 2 both a shock and rarefaction are attached to the trailing edge to turn the flow to the • horizontal

the flow regions are separated by a slipstream in which pressure and velocity directions • match

′ P P2 cos αo ( 2− ) F =(P ′ P ) A cos α ,C = 1 2 L 2 2 o L ρ1u • − 2 1

′ P P2 sin αo ( 2− ) FD =(P2′ P2) A sin αo,CD = 1 2 • − 2 ρ1u1 the drag here is known as wave drag or induced drag • other components of drag, skin friction drag and thickness drag are zero due to inviscid • limit and zero thickness limit

In the small disturbance limit

∆P γM 2 = 1 ∆θ (5.198) 2 P1 M1 1 −2 γP1M1 P ′ = P1 +p αo (5.199) 2 2 M1 1 2 − γP1M1 P2 = P1 + p ( αo) (5.200) 2 M1 1 − − 2 2γP1M1 P ′ Pp2 = αo (5.201) 2 2 − M1 1 − 2 2γPp1 u1 P2′ P2 = αo (5.202) − M 2 1 γP1 1 − ρ1 p 2 2 P2′ P2 = ρ1u1 αo (5.203) − M 2 1 1 − 2 2 FL = p ρ1u1 αoA cos αo (5.204) M 2 1 1 − 2 2 FLp= ρ1u1 αoA(1) (5.205) M 2 1 1 − 4αo p CL = (5.206) M 2 1 1 − CC BY-NC-ND.p 28 October 2019, J. M. Powers. 150 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

2 2 FD = ρ1u1 αoA sin αo (5.207) M 2 1 1 − 2 2 2 FpD = ρ1u α A (5.208) 2 1 o M1 1 − 2 4 αo p CD = (5.209) 2 M1 1 2− FL 2 M1 αo F L = = p (5.210) 2 2 ∗ ρ1c1A M1 1 2−2 FD 2 M1 αo F D = = p (5.211) 2 2 ∗ ρ1c1A M 1 1 − High Mach number limit: F =2M α (5.212) ∗Lp 1 o High Mach number limit: F =2M α2 (5.213) ∗D 1 o Dimensionless lift and drag are plotted versus Mach number in Figure 5.14

Example 5.3 Lift and Drag on an Inclined Flat Plate

Given: Flat plate, of chord length 2 m, depth 10 m inclined at 20◦ to the horizontal in a freestream of M1 = 3, P1 = 100 kPa, T1 = 300 K.

Find: Lift and drag forces on the plate.

Analysis: First some preliminaries:

J m c = γRT = (1.4) 287 (300 K) = 347.2 (5.214) 1 1 kg K s s   p m m u = M c = (3.0) 347.2 =1, 041.6 (5.215) 1 1 1 s s P1 100, 000 Pa  kg ρ1 = = =1.1614 3 (5.216) RT1 J m 287 kg K (300 K) γ   − γ 1 γ 1 P = P 1+ − M 2 (5.217) o 1 2 1   1 3.5 P = 100 kPa 1+ 32 = 367.327 kPa (5.218) o 5   In a previous example, we found the oblique shock state under identical conditions:

′ P2 = 377, 072 Pa (5.219)

Now consider the rarefaction.

CC BY-NC-ND. 28 October 2019, J. M. Powers. 5.7. SUPERSONIC FLOW OVER AIRFOILS 151

F* = F / ρ c 2 A L L 1 1 2 Dimensionless Lift Force 1.75 Flat Plate at Small Angle of Attack 1.5

1.25

1 Invalid Region

0.75

0.5 α ο 0.25 ο = 5

0 2 4 6 8 10 M 1 F* = F / ρ c 2 A D D 1 1 0.5

Dimensionless Drag Force 0.4 Flat Plate at Small Angle of Attack

0.3

0.2 Invalid Region

0.1 ο α ο = 5

0 2 4 6 8 10 M 1

Figure 5.14: Dimensionless Lift and Drag versus Incoming Mach Number for Flat Plate at Small Angle of Attack

γ +1 − γ 1 − ν(M )= tan 1 − (M 2 1) tan 1 M 2 1 (5.220) 1 γ 1 γ +1 1 − − 1 − r − r q 1.4+1 − 1.4 1 − ◦ ν(M )= tan 1 − (32 1) tan 1 32 1=0.8691 rad = 49.7973 (5.221) 1 1.4 1 1.4+1 − − − r − r ◦ p ν(M2)= ν(M1)+20 (5.222) ◦ ◦ ◦ ν(M2)=49.7973 + 20 = 69.7973 (5.223)

◦ 1.4+1 − 1.4 1 − 69.7973 =1.218 rad = tan 1 − (M 2 1) tan 1 M 2 1 (5.224) 1.4 1 1.4+1 2 − − 2 − r − r q M2 =4.3209 (5.225) − γ γ 1 γ−1 P = P 1+ − M 2 (5.226) 2 o 2 2  −  1 3.5 P = 367.327 kPa 1+ 4.32092 =1.591 kPa (5.227) 2 5   CC BY-NC-ND. 28 October 2019, J. M. Powers. 152 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

′ FL = (P2 P2) A cos αo (5.228) − ◦ F = (377, 072 Pa 1, 591 Pa)(10 m)(2 m)cos20 (5.229) L − FL =7, 142, 073 N (5.230) FL 7, 142, 073 N CL = = (5.231) 1 2 kg 2 ρ1u1A 1 m 2 2 1.1614 m3 1, 041.6 s (10 m)(2 m)

   CL =0.5668 (5.232) F = (P ′ P ) A sin α (5.233) D 2 − 2 o F = (377, 072 Pa 1, 591 Pa)(10 m) (2 m) sin 20◦ (5.234) D − FD =2, 320, 600 N (5.235) FD 2, 320, 600 N CD = = (5.236) 1 2 kg 2 ρ1u1A 1 m 2 2 1.1614 m3 1, 041.6 s (10 m)(2 m)

   CD =0.1842 (5.237) 4αo Compare with thin airfoil theory: CL thin = (5.238) M 2 1 1 − ◦ π rad 4(20 ) 180◦ p CL thin = =0.4936 (5.239) √32 1 − 2 4αo CD thin = (5.240) M 2 1 1 − ◦ π rad 2 4 (20 ) 180◦ p CD thin = =0.1723 (5.241) √32 1 − 

5.7.2 Diamond-shaped airfoil The simplest supersonic airfoil with camber for analysis purposes is the diamond shaped airfoil as sketched in Figure 5.15. The sketch shows the airfoil at zero angle of attack. The upper half of the wedge is inclined at angle ǫ to the horizontal In this case there will be no lift but there will be drag. Note the following features:

sudden turn through lead oblique shock • turn through isentropic Prandtl-Meyer rarefaction • final turn through oblique shock attached to trailing edge • far field limit: acoustic (Mach) waves • 4ǫ Thin airfoil limit CL thin = 2 , same as for flat plate! √M1 1 • − 4ǫ2 Thin airfoil limit CD thin = 2 , same as for flat plate! √M1 1 • −

CC BY-NC-ND. 28 October 2019, J. M. Powers. 5.7. SUPERSONIC FLOW OVER AIRFOILS 153

far field Mach waves

P1 T1 M1

P lead oblique 2 P3 shock P4 trailing oblique 2ε shock

Prandtl-Meyer rarefaction

Figure 5.15: Supersonic Flow over a Diamond-Shaped Airfoil

5.7.3 General curved airfoil A general airfoil with camber is sketched in Figure 5.16. The sketch shows the airfoil at zero angle of attack. In this case there will be no lift but there will be drag. Note the following features:

lead oblique shock • lead shock weakened by series of non-centered rarefaction waves • shock at trailing edge, also weakened by non-centered rarefaction waves • far field: acoustic (Mach) waves •

5.7.4 transition Transonic flow exists whenever there is a continuous transition from subsonic to supersonic flow. One example of a transonic flow is sketched in Figure 5.171 which shows an accelerating airfoil. Note:

1adopted from Bryson, A. E., “An Experimental Investigation of Transonic Flow Past Two-Dimensional Wedge and Circular-Arc Sections Using a Mach-Zehnder Interferometer,” NACA Tech. Note 2560, 1951.

CC BY-NC-ND. 28 October 2019, J. M. Powers. 154 CHAPTER 5. STEADY SUPERSONIC TWO-DIMENSIONAL FLOW

Figure 5.16: Supersonic Flow over a Curved Airfoil

Sonic Locus Sonic Locus Shock Shock M < 1 M > 1 M < 1 M < 1 M > 1 M < 1

Μ = 0.852 1 Μ1 = 0.892

M > 1 Shock Shock Shock

M > 1 M < 1 M > 1 M > 1 M < 1 Sonic Locus M > 1 M < 1

Μ1 = 1.207 Μ1 = 1.315 Μ1 = 1.465

Figure 5.17: Transition from Subsonic to Transonic to Supersonic Flow

for high subsonic Mach number a bubble of supersonic flow appears • smooth transition from subsonic to supersonic • shock transition from supersonic to subsonic • as Mach number increases, supersonic bubble expands • for slightly supersonic Mach number, new shock approaches from far field • as supersonic Mach number increases, shock from far field approaches leading edge and • supersonic bubble disappears

challenging problems, not easily solved till 1960’s! •

CC BY-NC-ND. 28 October 2019, J. M. Powers. Chapter 6

Linear flow analysis

see Anderson, Chapter 9

In this section we consider flows which are

steady, • two-dimensional, • irrotational, • isentropic, • calorically perfect, and • ideal. • The analysis is extensible to other cases.

6.1 Formulation

In lecture a detailed discussion is given in which the linearized velocity potential equation is obtained: ∂2φ ∂2φ 1 M 2 + =0. (6.1) − ∞ ∂x2 ∂y2  6.2 Subsonic flow

Here we consider flows in which the Mach number is subsonic, but not negligibly small.

155 156 CHAPTER 6. LINEAR FLOW ANALYSIS

6.2.1 Prandtl-Glauret rule A discussion is given where it is shown that the pressure coefficient on a supersonic airfoil can be determined in terms of the pressure coefficient known from subsonic theory:

cpo cp = . (6.2) 1 M 2 − ∞ p 6.2.2 Flow over wavy wall The technique of separation of variables is used to show the subsonic flow over a wavy wall can be written in terms of the velocity potential as

U h 2πx 2π 1 M 2 y φ(x, y)= ∞ sin exp − − ∞ . (6.3) 1 M 2 l l ! − ∞   p p 6.3 Supersonic flow

6.3.1 D’Alembert’s solution The D’Alembert solution for the wave equation is shown for supersonic flows:

φ(x, y)= f x + M 2 1y + g x M 2 1y . (6.4) ∞ − − ∞ −  p   p  6.3.2 Flow over wavy wall The solution for flow over a wavy wall is given in detail in lecture.

CC BY-NC-ND. 28 October 2019, J. M. Powers. Chapter 7

Viscous flow

This chapter will focus on problems in which viscous stress plays an important role in deter- mining the motion of the fluid. The topic in general is quite broad; to gain understanding of the fundamental physics, we will restrict our attention to the following limits:

incompressible fluid • isotropic Newtonian fluid with constant properties • at most two-dimensional unsteady flow • The chapter will consider the governing equations and then solve a few representative problems.

7.1 Governing equations

This section considers the governing equations for the conditions specified for this chapter. In dimensional non-conservative form, the governing equations are as follows:

∂u ∂v + =0 ∂x ∂y ∂u ∂u ∂u ∂P ∂2u ∂2u ρ + ρu + ρv = + µ + ∂t ∂x ∂y − ∂x ∂x2 ∂y2   ∂v ∂v ∂v ∂P ∂2v ∂2v ρ + ρu + ρv = + µ + ∂t ∂x ∂y − ∂y ∂x2 ∂y2   ∂T ∂T ∂T ∂P ∂P ∂P ∂2T ∂2T ρc + u + v = + u + v + k + p ∂t ∂x ∂y ∂t ∂x ∂y ∂x2 ∂y2       ∂u 2 1 ∂u ∂v 2 ∂v 2 +2µ + + + ∂x 2 ∂y ∂x ∂y       ! 157 158 CHAPTER 7. VISCOUS FLOW

An argument could be made to eliminate the viscous dissipation term and the pressure derivatives in the energy equation. The argument is subtle and based on the low Mach number limit which corresponds to incompressibility.

7.2 Couette flow

Consider a channel flow driven by plate motion See Figure 7.1

U T = T 0 y = h

P0 P0 y

x y = 0 x = 0 T = T 0

Figure 7.1: Sketch for Couette flow

The mechanics of such a flow can be described by stripping away many extraneous terms from the governing equations.

Take

fully developed velocity and temperature profiles: ∂u 0, ∂T 0 • ∂x ≡ ∂x ≡ steady flow ∂ 0 • ∂t ≡ constant pressure field P (x, y, t)= P • o constant temperature channel walls T (x, 0, t)= T (x, h, t)= T • o Since fully developed mass gives:

∂v = 0 (7.1) ∂y v(x, y)= f(x) (7.2) and since in order to prevent mass flowing through the wall boundaries, v(x, 0) = v(x, h) = 0, thus

CC BY-NC-ND. 28 October 2019, J. M. Powers. 7.2. COUETTE FLOW 159

v(x, y) = 0 (7.3)

∂u ∂u ∂T ∂T Since ∂x =0, ∂t = 0 and ∂x =0, ∂t = 0, we have at most,

u = u(y) (7.4) T = T (y) (7.5)

The y momentum equation has no information and x momentum and energy reduce to the following:

d2u 0= µ (7.6) dy2 d2T du 2 0= k + µ (7.7) dy2 dy   The x momentum equation is thus

d2u = 0 (7.8) dy2 du = C (7.9) dy 1

u(y)= C1y + C2 (7.10)

Now applying u(0) = 0 and u(h)= U to fix C1 and C2 we get

y u(y)= U (7.11) h Shear stress:

∂u τ = µ (7.12) yx ∂y U τ = µ (7.13) yx h The energy equation becomes

d2T µ du 2 = (7.14) dy2 −k dy   CC BY-NC-ND. 28 October 2019, J. M. Powers. 160 CHAPTER 7. VISCOUS FLOW

d2T µ U 2 = (7.15) dy2 −k h2 dT µ U 2 = y + C (7.16) dy −k h2 1 1 µ U 2 T (y)= y2 + C y + C (7.17) −2 k h2 1 2

Now T (0) = To and T (h)= To. This fixes the constants, so

1 µ y y 2 T (y)= U 2 + T (7.18) 2 k h − h o      In dimensionless form this becomes T T 1 µc U 2 y y 2 − o = p (7.19) T 2 k c T h − h o p o   T T Pr Ec  y   y 2 − o = (7.20) To 2 h − h µ      µc ν momentum diffusivity : Pr p = ρ = = (7.21) ≡ k k α thermal diffusivity ρcp U 2 kinetic energy : Ec = (7.22) ≡ cpTo thermal energy Now

dT 1 µ U 2 = (h 2y) (7.23) dy 2 k h2 − dT 1 U 2 q = k = µ (2y h) (7.24) y − dy 2 h2 − µU 2 q (0) = (7.25) y − 2h Note:

at lower wall, heat flux into wall; heat generated in fluid conducted to wall • wall heat flux magnitude independent of thermal conductivity • higher plate velocity, higher wall heat flux • higher viscosity, higher wall heat flux • thinner gap, higher wall heat flux •

CC BY-NC-ND. 28 October 2019, J. M. Powers. 7.3. SUDDENLY ACCELERATED FLAT PLATE 161

h Also since Tmax occurs at y = 2 1 µ T = U 2 + T (7.26) max 8 k o Note:

high viscosity, high maximum temperature • high plate velocity, high maximum temperature • low thermal conductivity, high maximum temperature • Dimensionless wall heat flux given by the :

qy(0) qy(0)∆y Nu k∆T = (7.27) ≡ k∆T ∆y µU 2 h

Nu = 2h 2 = 2 (7.28) 1 µ 2 k 8 k U

7.3 Suddenly accelerated flat plate

The problem of pulling a plate suddenly in a fluid which is initially at rest is often known as Stokes’ First Problem or Rayleigh’s problem.

7.3.1 Formulation Consider a channel flow driven by a suddenly accelerated plate. See Figure 7.2 Initially, t< 0

fluid at rest • plate at rest • For t 0 ≥ plate pulled at constant velocity U • Assume:

constant pressure P (x, y, t)= P • o fully developed flow ∂u =0, ∂T =0 • ∂x ∂x

CC BY-NC-ND. 28 October 2019, J. M. Powers. 162 CHAPTER 7. VISCOUS FLOW

u( ∞ , t) = 0 ∞ T( , t) = T0 u (y, 0) = 0

T(y, t) = T0 ρ, µ, P0

u = u (y, t) 1/2 y δ ∼ (µ t / ρ)

x u(0,t) = U U T(0, t) = T0

Figure 7.2: Sketch for Stokes’ First Problem

Again from mass we deduce that v(x, y, t) = 0. The x momentum equation reduces to

∂u ∂2u ρ = µ (7.29) ∂t ∂y2 The initial and boundary conditions are

u(y, 0) = 0 (7.30) u(0, t)= U (7.31) u( , t) = 0 (7.32) ∞ 7.3.2 Velocity profile This problem is solved in detail in lecture. The solution for the velocity field is shown to be

u 2 y/√νt =1 exp s2 ds (7.33) U − √π − o Z0  7.4 Starting transient for plane Couette flow

The starting transient problem for plane Couette flow can be formulated as

∂u ∂2u = ν (7.34) ∂t ∂y2

u(y, 0)=0, u(0, t)= Uo, u(h, t)=0. (7.35) In class a detailed solution is presented via the technique of separation of variables. The solution is

CC BY-NC-ND. 28 October 2019, J. M. Powers. 7.5. BLASIUS BOUNDARY LAYER 163

u y 2 ∞ 1 n2π2νt nπy =1 exp − sin (7.36) U − h − π n h2 h o n=1 X     7.5 Blasius boundary layer

The problem of flow over a flat plate in the absence of pressure gradient is formulated and solved using the classical approach of Blasius.

7.5.1 Formulation After suitable scaling and definition of similarity variables, discussed in detail in class, the following third order non-linear ordinary differential equation is obtained:

d3f 1 d2f + f =0, (7.37) dη3 2 dη2 df df =0, =1, f =0. (7.38) dη dη |η=0 η=0 η →∞

This equation is solved numerically as a homework problem.

7.5.2 Wall shear stress The solution is used to obtain the classical formulae for skin friction coefficient: 0.664 Cf = , (7.39) √Rex and drag coefficient: 1.328 CD = . (7.40) √ReL

CC BY-NC-ND. 28 October 2019, J. M. Powers. 164 CHAPTER 7. VISCOUS FLOW

CC BY-NC-ND. 28 October 2019, J. M. Powers. Chapter 8

Acoustics

This chapter outlines the brief introduction to acoustics given in class in somewhat more detail.

8.1 Formulation

We reduce the Euler equations for isentropic flow to the following equations where quantities with a hat are understood to be small perturbations about the ambient state, denoted with a subscript of ”o”.

∂ρˆ + ρ vˆ = 0 (8.1) ∂t o∇· ∂vˆ ρ + Pˆ = 0 (8.2) o ∂t ∇· ˆ 2 P = coρ.ˆ (8.3)

Introducing the velocity potential φ = vˆ and employing further manipulation allows the equation to be written as the wave∇ equation:

∂2φ = c2 2φ. (8.4) ∂t2 o∇

The pressure, velocity, and density are then obtained from

∂φ Pˆ = ρ , (8.5) − o ∂t vˆ = φ, (8.6) ∇ ∂φ ρˆ = ρ c2 . (8.7) − o o ∂t

165 166 CHAPTER 8. ACOUSTICS

8.2 Planar waves

The D’Alembert solution for planar waves is shown in class to be

φ = f(x + c t)+ g(x c t), (8.8) o − o Pˆ = ρ c f ′(x + c t)+ ρ c g′(x c t), (8.9) − o o o o o − o uˆ = f ′(x + c t)+ g′(x c t), (8.10) o − o (8.11)

8.3 Spherical waves

The D’Alembert solution for spherical waves is shown in class to be 1 1 φ = f(r + c t)+ g(r c t), (8.12) r o r − o ρoco ρoco Pˆ = f ′(r + c t)+ g′(r c t), (8.13) − r o r − o 1 1 uˆ = f ′(r + c t)+ g′(r c t), (8.14) r o r − o (8.15)

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