You are on page 1of 184

The Pennsylvania State University

The Graduate School


NUMERICAL SIMULATIONS OF ACOUSTICS PROBLEMS
USING THE DIRECT SIMULATION MONTE CARLO METHOD
A Dissertation in
Acoustics
by
Amanda Danforth Hanford
c _ 2008 Amanda Danforth Hanford
Submitted in Partial Fulllment
of the Requirements
for the Degree of
Doctor of Philosophy
August 2008
The Dissertation of Amanda Danforth Hanford was reviewed and approved

by
the following:
Lyle N. Long
Distinguished Professor of Aerospace Engineering and Acoustics
Dissertation Advisor, Chair of Committee
James B. Anderson
Evan Pugh Professor of Chemistry and Physics
Feri Farassat
Senior Theoretical Aeroacoustician, NASA Langley Research Center
Special Member
Thomas B. Gabrielson
Professor of Acoustics
Victor W. Sparrow
Professor of Acoustics
Anthony A. Atchley
Professor of Acoustics
Chair of Graduate Program in Acoustics

Signatures are on le in the Graduate School.


Abstract
In the current study, real gas eects in the propagation of sound waves are simu-
lated using the direct simulation Monte Carlo method for a wide range of systems.
This particle method allows for treatment of acoustic phenomena for a wide range
of Knudsen numbers, dened as the ratio of molecular mean free path to wave-
length. Continuum models such as the Euler and Navier-Stokes equations break
down for ows greater than a Knudsen number of approximately 0.05. Continuum
models also suer from the inability to simultaneously model nonequilibrium con-
ditions, diatomic or polyatomic molecules, nonlinearity and relaxation eects and
are limited in their range of validity. Therefore, direct simulation Monte Carlo
is capable of directly simulating acoustic waves with a level of detail not possible
with continuum approaches.
The basis of direct simulation Monte Carlo lies within kinetic theory where
representative particles are followed as they move and collide with other parti-
cles. A parallel, object-oriented DSMC solver was developed for this problem.
Despite excellent parallel eciency, computation time is considerable. Monatomic
gases, gases with internal energy, planetary environments, and amplitude eects
spanning a large range of Knudsen number have all been modeled with the same
method and compared to existing theory. With the direct simulation method,
signicant deviations from continuum predictions are observed for high Knudsen
number ows.
iii
Table of Contents
List of Figures viii
List of Tables xiii
List of Symbols xiv
Acknowledgments xviii
Chapter 1
Introduction 1
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Numerical models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 Continuum methods . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.4 Particle methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.5 Sound for all Knudsen numbers . . . . . . . . . . . . . . . . . . . . 6
1.6 Direct Simulation Monte Carlo . . . . . . . . . . . . . . . . . . . . 7
1.6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
Chapter 2
Kinetic Theory of Gases 9
2.1 Historical background and
Ludwig Boltzmann . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Velocity distribution function . . . . . . . . . . . . . . . . . . . . . 11
2.2.1 Macroscopic properties in a simple gas . . . . . . . . . . . . 11
2.3 Deriving the Boltzmann equation . . . . . . . . . . . . . . . . . . . 16
2.3.1 The calculation of the collision integral . . . . . . . . . . . . 17
2.4 Deriving conservation equations
from the Boltzmann equation . . . . . . . . . . . . . . . . . . . . . 19
iv
2.5 Solutions to the Boltzmann equation . . . . . . . . . . . . . . . . . 22
2.5.1 Equilibrium properties . . . . . . . . . . . . . . . . . . . . . 22
2.5.2 Linearized Boltzmann Equation . . . . . . . . . . . . . . . . 24
2.5.2.1 1st order solution . . . . . . . . . . . . . . . . . . . 25
2.5.2.2 Transport coecients . . . . . . . . . . . . . . . . 26
2.5.3 Bhatnagar, Gross, and Krook (BGK) equation . . . . . . . . 28
2.5.4 Numerical solutions to the Boltzmann equation . . . . . . . 29
Chapter 3
Implementation of DSMC for Acoustics Simulations 31
3.1 Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.1.1 Initialization . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.1.2 Move particles and Boundary Conditions . . . . . . . . . . . 33
3.1.3 Sort particles . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.1.4 Collision . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.1.4.1 Binary collisions . . . . . . . . . . . . . . . . . . . 35
3.1.4.2 DSMC collision routine . . . . . . . . . . . . . . . 37
3.1.5 Sampling . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.2 Diatomic and polyatomic gases . . . . . . . . . . . . . . . . . . . . 40
3.2.1 DSMC tests for gases with internal energy . . . . . . . . . . 43
3.3 Mixtures implementation . . . . . . . . . . . . . . . . . . . . . . . . 44
3.4 Assumptions and Error within DSMC . . . . . . . . . . . . . . . . . 45
Chapter 4
Computing Issues 50
4.1 Object-Oriented Programming Approach . . . . . . . . . . . . . . . 50
4.2 Parallel Implementation . . . . . . . . . . . . . . . . . . . . . . . . 53
4.3 Modications to the DSMC algorithm . . . . . . . . . . . . . . . . . 58
4.3.1 Small deviation from equilibrium . . . . . . . . . . . . . . . 58
4.3.2 Low Kn ows modications . . . . . . . . . . . . . . . . . . 60
Chapter 5
Absorption and Dispersion in a Monatomic Gas 62
5.1 The speed of sound . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.2 The absorption of sound . . . . . . . . . . . . . . . . . . . . . . . . 63
5.3 Theory on the absorption and
dispersion of sound in a monatomic gas . . . . . . . . . . . . . . . . 64
5.4 DSMC Results for the absorption
and dispersion of sound in a monatomic gas . . . . . . . . . . . . . 69
5.4.1 Simulation approach . . . . . . . . . . . . . . . . . . . . . . 69
v
5.4.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
5.4.2.1 Nonequilibrium . . . . . . . . . . . . . . . . . . . . 72
Chapter 6
Absorption and Dispersion in a Gas With Internal Energy 75
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
6.2 Absorption and dispersion from a simple
relaxation process . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
6.3 Current theories including rotational
relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
6.3.1 Other theories for rotational relaxation . . . . . . . . . . . . 80
6.4 Current theories including vibrational
relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
6.5 Implementation in DSMC for a gas with
rotational energy for multiple collision
numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
6.5.1 Simulation approach . . . . . . . . . . . . . . . . . . . . . . 83
6.5.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
6.6 Implementation in DSMC for a gas with
rotational and vibrational energy . . . . . . . . . . . . . . . . . . . 87
6.6.1 Simulation approach . . . . . . . . . . . . . . . . . . . . . . 87
6.6.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
6.6.2.1 Nonequilibrium . . . . . . . . . . . . . . . . . . . . 90
6.6.2.2 Absorption as a function of temperature . . . . . . 92
6.6.2.3 Dispersion as a function of temperature . . . . . . 95
Chapter 7
The Eect of Amplitude 98
7.1 Simulation approach . . . . . . . . . . . . . . . . . . . . . . . . . . 98
7.2 Breakdown of the propagation constant . . . . . . . . . . . . . . . . 99
7.3 Nonequilibrium eects as a function
of amplitude . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
7.4 Absorption and dispersion as a function of amplitude . . . . . . . . 105
7.5 Harmonic generation . . . . . . . . . . . . . . . . . . . . . . . . . . 109
7.6 Shock coalescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
7.6.1 Coalescence at high Kn . . . . . . . . . . . . . . . . . . . . 116
Chapter 8
DSMC Applications: Planetary Acoustics 119
8.1 Simulation approach . . . . . . . . . . . . . . . . . . . . . . . . . . 121
vi
8.2 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . 121
8.2.1 Earth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
8.2.2 Mars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
8.2.3 Titan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
8.2.4 Vertical proles . . . . . . . . . . . . . . . . . . . . . . . . . 131
Chapter 9
Conclusions 137
9.1 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
Appendix A
Buons Needle Experiment 142
Appendix B
Sampling a Maxwellian Distribution 144
Appendix C
Deriving the Navier-Stokes Dispersion Relation 147
Bibliography 150
vii
List of Figures
1.1 Limits of applicability of various mathematical models to simulate
uid ow [1] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1 Ludwig Boltzmann (1844-1906) . . . . . . . . . . . . . . . . . . . . 10
2.2 Maxwellian velocity distribution function for argon at 0

. . . . . . 23
3.1 A owchart of the DSMC algorithm. There are NST total samples
with NIS time steps in between samples. J and I increment NST
and NIS respectively. . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.2 A owchart of the collision routine in the DSMC algorithm. sigV-
max is calculated by looping through all the particles in the cell,
the number of collisions to be performed, numColl, is given by Eq.
(3.15) and the probability of collision, prob, is given by Eq. (3.14).
Details of the internal energy exchange routine will be given in Sec.
(3.2). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.3 A owchart of the internal energy exchange routine in the DSMC
algorithm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.4 DSMC simulation of nitrogen molecules undergoing classical relax-
ation at 4000 K. Birds exponential model[1] given by solid lines
and DSMC results in dashed lines [2]. . . . . . . . . . . . . . . . . . 44
3.5 DSMC simulation of nitrogen molecules undergoing relaxation with
a coupled discrete vibration / classical rotation model at 4000 K.
Birds exponential model[1] given by solid lines and DSMC results
in dashed lines [2]. . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.6 Equilibrium initialization of argon at 0

C after 1 ensemble . . . . . 47
3.7 Equilibrium initialization of argon at 0

C after 10 ensembles . . . . 47
3.8 Equilibrium initialization of argon at 0

C after 100 ensembles . . . 48


3.9 Equilibrium initialization of argon at 0

C after 1000 ensembles . . 48


4.1 A owchart of the parallel algorithm . . . . . . . . . . . . . . . . . 54
viii
4.2 CPU time on Columbia and Mufasa with respect to number of
processors compared to ideal CPU time . . . . . . . . . . . . . . . . 56
4.3 Speedup on Columbia and Mufasa with respect to the number of
processors compared to ideal speedup . . . . . . . . . . . . . . . . . 57
4.4 Parallel eciency on Columbia and Mufasa with respect to the
number of processors . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.5 Parallel eciency on Mufasa with respect to the number of proces-
sors for a large and small system . . . . . . . . . . . . . . . . . . . 59
5.1 Scaled absorption
cl
/k
0
and dispersion
cl
/k
0
predictions given by
the linearized Navier-Stokes equations from Eq. (5.20) are plotted
with the low frequency classical absorption coecient given by Eq.
(5.21) [3] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.2 Scaled absorption
cl
/k
0
in argon for 273 K. DSMC results (red
circles) compared to experimental results by Greenspan [4] (green
triangles) and Schotter[5] (blue squares) and continuum theory pre-
dictions results from Eq. (5.20) [3] . . . . . . . . . . . . . . . . . . 71
5.3 Scaled dispersion k/k
0
in argon for 273 K. DSMC results (red cir-
cles) compared to experimental results by Greenspan [4] (green tri-
angles) and continuum theory predictions results from Eq. (5.20)
[3] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
5.4 Translational nonequilibrium eects for Kn = 2.0 at 273 K in argon. 73
5.5 Translational nonequilibrium eects for Kn = 0.02 at 273 K in argon. 74
6.1 Scaled absorption for collision numbers Z
rot
=1, 10, 100, and 1000
in a nitrogen-like gas at 0

C from Eqs. (6.7) and (6.3) [3, 6] . . . . 79


6.2 Scaled dispersion for collision numbers Z
rot
=1, 10, 100, and 1000
in a nitrogen-like gas at 0

C from Eqs. (6.8) and (6.2) [3, 6] . . . . 79


6.3 Vibrational relaxation frequency for oxygen and nitrogen as a func-
tion of temperature [7, 8] . . . . . . . . . . . . . . . . . . . . . . . 82
6.4 Scaled absorption for relaxation collision number of 1. DSMC simu-
lations (points) are plotted with continuum theory for the rotational
relaxation given by Eq. (6.3) (dashed line) [3] and total absorption
given by Eq. (6.7) (solid line) [6] . . . . . . . . . . . . . . . . . . . 85
6.5 Scaled absorption for relaxation collision number of 5. DSMC simu-
lations (points) are plotted with continuum theory for the rotational
relaxation given by Eq. (6.3) (dashed line) [3] and total absorption
given by Eq. (6.7) (solid line) [6] . . . . . . . . . . . . . . . . . . . 85
ix
6.6 Scaled absorption for relaxation collision number of 40. DSMC
simulations (points) are plotted with continuum theory for the ro-
tational relaxation given by Eq. (6.3) (dashed line) [3] and total
absorption given by Eq. (6.7) (solid line) [6] . . . . . . . . . . . . . 86
6.7 Scaled absorption for relaxation collision number of 200. DSMC
simulations (points) are plotted with continuum theory for the ro-
tational relaxation given by Eq. (6.3) (dashed line) [3] and total
absorption given by Eq. (6.7) (solid line) [6] . . . . . . . . . . . . . 86
6.8 DSMC results for the scaled absorption with relaxation collision
numbers of 1, 5, 40, and 200. DSMC simulations (points) are plot-
ted with continuum theory for rotational relaxation (dashed line)
[3] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
6.9 Nonequilibrium eects for Kn = 0.02 at 273 K with classical vibra-
tion model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
6.10 Nonequilibrium eects for Kn = 1.0 at 273 K with classical vibra-
tion model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
6.11 Fraction of molecules in the excited state for Kn = 0.02 at 273 K
(red square), 2000 K (green triangle) and 4000 K (blue circle). . . . 92
6.12 (Scaled absorption in nitrogen for 273 K. DSMC results (symbols)
compared to continuum theory predictions given by Eqs. (5.20),
(5.21), and (6.7) (lines) [3, 6]. . . . . . . . . . . . . . . . . . . . . . 93
6.13 Scaled absorption in nitrogen for 2000 K. DSMC results (symbols)
compared to continuum theory predictions given by Eqs. (5.20),
(5.21), and (6.7) (lines) [3, 6]. . . . . . . . . . . . . . . . . . . . . . 94
6.14 Scaled absorption in nitrogen for 4000 K. DSMC results (symbols)
compared to continuum theory predictions given by Eqs. (5.20),
(5.21), and (6.7) (lines) [3, 6]. . . . . . . . . . . . . . . . . . . . . . 95
6.15 Scaled dispersion in nitrogen for 273 K. DSMC results (symbols)
compared to continuum theory predictions given by Eqs. (5.20),
(6.2), and (6.8) (lines) [3, 6]. . . . . . . . . . . . . . . . . . . . . . . 96
6.16 Scaled dispersion in nitrogen for 2000 K. DSMC results (symbols)
compared to continuum theory predictions given by Eqs. (5.20),
(6.2), and (6.8) (lines) [3, 6]. . . . . . . . . . . . . . . . . . . . . . . 97
6.17 Scaled dispersion in nitrogen for 4000 K. DSMC results (symbols)
compared to continuum theory predictions given by Eqs. (5.20),
(6.2), and (6.8) (lines) [3, 6]. . . . . . . . . . . . . . . . . . . . . . . 97
7.1 Maximum pressure amplitude (Pa) in argon for Kn = 2, 40 m/s
amplitude. DSMC results (points) compared to restricted exponen-
tial best t curve for x < 10u

/ (solid line) . . . . . . . . . . . . . 100


x
7.2 Maximum pressure amplitude (Pa) in argon for Kn = 2, 40 m/s
amplitude. DSMC results (points) compared to exponential best
t curve for x > 10u

/ (solid line) . . . . . . . . . . . . . . . . . . 101


7.3 Maximum pressure amplitude (Pa) in argon for Kn = 2, 5 m/s
amplitude. DSMC results (points) compared to exponential best
t curve for x > 10u

/ (solid line) . . . . . . . . . . . . . . . . . . 101


7.4 Maximum pressure amplitude (Pa) in argon for Kn = 0.02, 40 m/s
amplitude. DSMC results (points) compared to exponential best
t curve for x > 10u

/ (solid line) . . . . . . . . . . . . . . . . . . 102


7.5 T
trx
/T
tr
for Kn = 2, 40 m/s amplitude. . . . . . . . . . . . . . . . . 103
7.6 T
trx
/T
tr
for Kn = 2, 5 m/s amplitude. . . . . . . . . . . . . . . . . 104
7.7 T
trx
/T
tr
for Kn = 0.02, 40 m/s amplitude. . . . . . . . . . . . . . . 104
7.8 T
trx
/T
tr
for Kn = 0.02, 5 m/s amplitude. . . . . . . . . . . . . . . . 105
7.9 Amplitude dependence on the scaled absorption in argon at 0

as
a function of Kn. DSMC results (points) compared to continuum
theory given by Eq. (5.20) (line) . . . . . . . . . . . . . . . . . . . . 106
7.10 Wave steepening at Kn = 0.02 and 40 m/s amplitude . . . . . . . . 107
7.11 Amplitude dependence on the scaled dispersion in argon at 0

as a
function of Kn based on the maximum pressure amplitude. DSMC
results (points) compared to continuum theory given by Eq. (5.20)
(line) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
7.12 Amplitude dependence on the scaled dispersion in argon at 0

as a
function of Kn based on the zero crossings of the acoustic pressure.
DSMC results (points) compared to continuum theory given by Eq.
(5.20) (line) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
7.13 Number of collisions performed in each cell for a 40 m/s amplitude
(dashed line) and 5 m/s amplitude (solid line) at Kn = 0.2 after a
time of 1 nanosecond . . . . . . . . . . . . . . . . . . . . . . . . . . 110
7.14 Fourier component amplitude for Kn = 0.005, 40 m/s amplitude
compared to Fubini solution given by Eq. (7.4) and Burgers equa-
tion given by Eq. (7.5) [9, 10]. . . . . . . . . . . . . . . . . . . . . . 112
7.15 Fourier component amplitude for Kn = 2, 40 m/s amplitude. . . . . 113
7.16 Fourier component amplitude for Kn = 2, 5 m/s amplitude. . . . . 113
7.17 Shock coalescence for Kn = 0.02 at 6 nanoseconds . . . . . . . . . . 114
7.18 Shock coalescence for Kn = 0.02 at 15 nanoseconds . . . . . . . . . 115
7.19 Shock coalescence for Kn = 0.02 at 21 nanoseconds . . . . . . . . . 115
7.20 Shock coalescence for Kn = 0.02 at 46 nanoseconds . . . . . . . . . 116
7.21 Coalescence for Kn = 2 at 42 nanoseconds . . . . . . . . . . . . . . 117
7.22 Coalescence for Kn = 2 at 63 nanoseconds . . . . . . . . . . . . . . 117
7.23 Coalescence for Kn = 2 at 74 nanoseconds . . . . . . . . . . . . . . 118
xi
7.24 Coalescence for Kn = 2 at 84 nanoseconds . . . . . . . . . . . . . . 118
8.1 Temperature and pressure proles as a function of altitude above
Earths surface [11] . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
8.2 Scaled absorption for dry air on Earth. DSMC simulations (points)
are plotted with continuum theory for the vibrational relaxation
given by Eq. (6.3) (green) and total absorption given by Eq. (6.7)
(red) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
8.3 Scaled absorption for humid air on Earth. DSMC simulations
(points) are plotted with continuum theory for the vibrational re-
laxation given by Eq. (6.3) (green) and total absorption given by
Eq. (6.7) (red) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
8.4 DSMC results (solid line) for the acoustic pressure on Earth for
Kn = 0.02 compared to predicted amplitude dependence deter-
mined from the Navier-Stokes derived absorption coecient from
Eq. (5.20)(dashed line) . . . . . . . . . . . . . . . . . . . . . . . . . 125
8.5 Temperature and pressure proles as a function of altitude above
the Mars surface [12] . . . . . . . . . . . . . . . . . . . . . . . . . . 126
8.6 DSMC results for the acoustic pressure amplitude as a function of
distance on Mars for Kn = 0.02 compared to theoretical predictions
[13] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
8.7 Nonequilibrium eects on Mars showing rotational (dashed line)
and translational (solid line) temperatures for Kn = 2 . . . . . . . . 128
8.8 Temperature and pressure proles as a function of altitude above
Titans surface [14] . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
8.9 Kn = 0.042 waveform on Earth with = 0.14. Absorption domi-
nates nonlinearity with little or no nonlinear eects visible. . . . . . 131
8.10 Kn = 0.01 waveform on Titan with = 0.14. Signicant wave
steepening can be observed. . . . . . . . . . . . . . . . . . . . . . . 132
8.11 Knudsen number as a function of altitude on Earth for frequencies
of 100 Hz, 1000 Hz, 10000 Hz, and 1000000 Hz . . . . . . . . . . . . 133
8.12 Knudsen number as a function of altitude on Mars for frequencies
of 100 Hz, 1000 Hz, 10000 Hz, and 1000000 Hz . . . . . . . . . . . . 134
8.13 Knudsen number as a function of altitude on Titan for frequencies
of 100 Hz, 1000 Hz, 10000 Hz, and 1000000 Hz . . . . . . . . . . . . 135
8.14 The scaled absorption for a 70 MHz signal on Earth (blue), Mars
(red), and Titan (green) at an altitude of 25 km compared to Navier-
Stokes predicted thermal-viscous losses (black line) . . . . . . . . . 136
A.1 An illustration of the Buons needle experiment. . . . . . . . . . . 142
xii
List of Tables
6.1 The Prandlt number for common gases . . . . . . . . . . . . . . . . 78
8.1 Atmospheric conditions at the surface on Earth, Mars and Titan . . 120
xiii
List of Symbols
c Speed of sound, phase speed, [m/s], p. 63
c
p
Specic heat at constant pressure, [J/(kg K)], p. 16
C
p
Heat capacity at constant pressure, [J/(kmol K)], p. 15
c
v
Specic heat at constant volume, [J/(kg K)], p. 16
C
v
Heat capacity at constant volume, [J/(kmol K)], p. 15
c
0
Low amplitude, low frequency speed of sound, [m/s], p. 63
C

Heat capacity of relaxing mode, [J/(kmol K)], p. 76


d Distance, [m], p. 142
e
int
Internal energy per unit mass, [J/kg], p. 14
f Velocity distribution function, p. 11
f Frequency, [Hz], p. 76
F Force eld per unit mass, p. 16
F
num
Ratio of real to simulated particles, p. 37
f
0
Maxwellian velocity distribution function, p. 23
f
r
Relaxation frequency, [Hz], p. 76
J Bessel function, p. 110
k Complex propagation constant, [1/m], p. 64
xiv
k
b
Boltzmanns constant, [J/K], p. 13
k
cl
Complex propagation constant for classical losses, [1/m], p. 64
k
0
Wave number, [1/m], p. 67
Kn Knudsen number, p. 1
m Molecular mass, [kg/molecule], 11
M Molecular weight, [kg/kmol], 14
n Number density, [molecules/m
3
], p. 11
^ Avogadros number, [molecules/kmol], p. 13
N
c
Number of molecules in a cell, p. 37
p Pressure tensor, [Pa], p. 13
p Hydrostatic pressure, [Pa], p. 13
q Heat ux vector, [W/m
2
], p. 15
Q Macroscopic quantity, p. 12
r Particle position vector, [m], p. 11
1 Universal gas constant, [J/(kmol K)], p. 13
s Relaxation strength, p. 76
S Sonine polynomials, p. 25
t Time, [s], p. 11
T Temperature, [K], p. 13
T
int
Internal temperature, [K], p. 14
T
tr
Translational temperature, [K], p. 14
T
ov
Overall temperature, [K], p. 14
u Molecular velocity, [m/s], p. 11
U Mean velocity, [m/s], p. 12
xv
U
0
Macroscopic velocity amplitude, [m/s], p. 107
u

Thermal velocity, [m/s], p. 12


u

Average thermal speed, [m/s], p. 23


u

m
Most probable molecular thermal speed, [m/s], p. 23
u
r
Relative speed, [m/s], p. 18
x Shock formation distance, [m], p. 109
Z Relaxation collision number, p. 40
Z
rot
Rotational relaxation collision number, p. 77
Z
vib
Vibrational relaxation collision number, p. 87
Absorption coecient, [np/m], p. 64

cr
Combined absorption due to classical thermal-viscous losses and rotational
relaxation, [np/m], p. 77

r
Absorption due to a single relaxation process, [np/m], p. 76

rot
Absorption due to rotational relaxation, [np/m], p. 77

vib
Absorption due to vibrational relaxation, [np/m], p. 82
Dispersion coecient, [1/m], p. 64

cr
Combined dispersion due to classical thermal-viscous losses and rotational
relaxation, [1/m], p. 77

NL
Coecient of nonlinearity, p. 107

m
Reciprocal of the most probable molecular thermal speed, [s/m], p. 23

r
Dispersion due to a single relaxation process, [1/m], p. 76

rot
Dispersion due to rotational relaxation, [1/m], p. 77
Ratio of specic heats, p. 15
Goldberg number, p. 111
Kronecker delta function, p. 13
xvi
Acoustic Mach number, p. 109
Number of internal degrees of freedom, p. 14
Thermal conductivity, [J/K m s)], p. 26

m
Mean free path, [m], p. 1
Coecient of viscosity, [kg/(m s)], p. 26

b
Bulk viscosity, [kg/(m s)], p. 27
Collision frequency, [1/s], p. 76
Number of total degrees of freedom, p. 15
Mass density, [kg/m
3
], p. 11
Collision cross section, [kg m
2
], p. 17

NL
Nondimensional shock formation distance, p. 110
Viscous stress tensor, [Pa], p. 13

r
Relaxation time, [1/s], p. 76
Macroscopic quantity, p. 64
Perturbation from distribution function, p. 24
Dissipation function, p. 21
Frequency, [rad/s], p. 64
Scattering angle, p. 17
xvii
Acknowledgments
I would like to give my sincere thanks to Prof. Lyle N. Long for his advice and
support throughout my pursuit of this degree. This project would not have been
possible without his guidance.
I express my gratitude to my other committee members, Dr. James Anderson,
Dr. Thomas Gabrielson, Dr. Victor Sparrow, and Dr. Feri Farassat for their
comments and suggestions.
I would also like to acknowledge the National Science Foundation for fund-
ing the Consortium for Education in Many-Body Applications (CEMBA), Grant
No. NSF-DGE-9987589, and the NASA Graduate Student Fellowship Program for
funding and providing computer resources for this research project.
A huge thanks to Dr. Patrick D. OConnor. Ive said it before, but Ill say it
again, if Im getting a PhD, he should be getting two for all the hard work and
countless hours of conversation weve had over the years. I honestly wouldnt be
where I am today without his help and friendship.
Special thanks to Rebecca Sanford DeRousie and Bernadette Rakszawski for
making the littleman laugh while I typed away and to Catherine Hofstetter whose
friendship and motherhood I will always admire. And a huge thanks to Jen
Marcovich, Mahreen George, Jen Dombroskie, Suzi Lang, Shelley Farahani, Julie
Willits, Allison Bohn and Bethany Heim for being such a loving and enriching
part of our lives and whose support and friendship have meant so much to me.
And thanks also to the La Leche League of State College and Centre County
Babywearers for enriching our lives and making me the mother I am today.
Of course, I also owe a large debt of gratitude to my family. My parents,
Rhoda and Larry Danforth have given me countless words of encouragement and
support over the years. To my sister and her family Becky, Hans, Chloe, Paige,
and Peanut Watz for plenty of phone calls and stories to help me through the
day. To my brother, David Danforth for always keeping me on my toes with love
and big hugs. And much love and thanks to Pat, Keith and Kim Hanford for all
of the loving, warm encouragement and support in all things. Many thanks to
xviii
Karl, Melissa, Alice and Liz Sweitzer for being such an amazing part of my life,
giving me rides, talking Matlab, making cookies and hair dying. Thank you all for
making me feel like the smartest human alive, even when my day to day activities
prove that I am not.
I am indebted to my husband, Scott, for providing me motivation and encour-
agement at every step during this pursuit, for being with me through thick and
thin, and for the joy and happiness his company brings me. And many thanks to
Noah, the best littleman a mother could have, for reminding me that unconditional
love triumphs any degree.
xix
Dedication
To mothers and children everywhere. Children can teach you so much more about
life than books.
xx
Chapter 1
Introduction
1.1 Motivation
Although popular in acoustics, continuum models such as the Euler and Navier-
Stokes equations break down for ows approaching the well-known continuum limit
dened by the Knudsen number. Continuum models also suer from the inabil-
ity to simultaneously model nonequilibrium conditions, diatomic or polyatomic
molecules, nonlinearity and relaxation eects and are limited in their range of va-
lidity. Theory for the behavior of sound at high Knudsen numbers is inconsistent
with experimental results and needs to be resolved. Therefore, a single method
that spans the entire Knudsen number range is desirable. The best approach,
one of the few available, is to use a particle-based model based on the Boltzmann
equation to bypass continuum limitations while also minimizing the numerical com-
plexity of the model itself. Therefore, a parallel, object-oriented direct simulation
Monte Carlo (DSMC) method was chosen for this study in the application to the
simulation of acoustic wave propagation.
1.2 Numerical models
Fluid dynamics models for a gas can be categorized into two groups: continuum
methods and particle methods. Continuum methods, which are widely used for
acoustic problems, model the uid as a continuous medium. This model describes
2
the state of the uid macroscopically using quantities such as density, pressure,
velocity, and temperature. The continuum approximation is valid when the char-
acteristic length of the problem is much larger than the molecular mean free path
(
m
). The Knudsen number (Kn) is a nondimensional parameter dened as the
mean free path divided by a characteristic length and is a measure of the thermal
nonequilibrium in a gas [15]. The continuum condition (Kn < 0.05) is satised
for many engineering problems, which can be described using continuum equations
such as the Navier-Stokes or Euler equations.
Particle methods are based on molecular models that describe the state of the
gas at the microscopic level. The mathematical model at this level is the Boltzmann
equation which will be derived and discussed in detail in Chapter (2). Despite the
fact that the Boltzmann equation has been derived using a microscopic approach
it will also been shown in Chapter (2) that the Boltzmann equation will reduce to
the continuum conservation equations (e.g., Navier-Stokes) for low Kn [1, 15].
1.3 Continuum methods
The classical linear theory of sound propagation assumes that all acoustic elds
can be written as the sum of an equilibrium value
0
and a small perturbation
from equilibrium

. The governing continuum equations that describe the per-


turbation from equilibrium

can be derived from linearizing the Navier-Stokes


or Euler equations. During the linearization process, higher order terms are ig-
nored by assuming the perturbation from equilibrium

is small. Many valuable


acoustic phenomenon can be developed under this assumption (see reference [9]).
However, using the linearized versions of the continuum equations is only valid for
small acoustic amplitudes. Computational methods in linear acoustics are fast and
simple but the restriction to low acoustic amplitudes is undesirable.
Keeping higher order terms allows for exploration in nonlinear acoustics where

does not necessarily have to be small to describe nonlinear phenomena such


as shock wave formation and harmonic generation [10]. An overview of compu-
tational approaches in nonlinear acoustics reveals that nonlinear eects are often
treated separately using the nonlinear Euler equations or Burgers equation, and
then superimposed to produce a nal result, often going back and forth between
3
the time and frequency domains [16, 17]. These methods are often tailored for quite
specic cases, and usually model only one-dimensional propagation, hindering the
general utility of such models. A method by Sparrow and Raspet [18] that con-
tains up to second-order nonlinear terms provides a multidimensional simulation
tool for nonlinear acoustics, but the method did not explicitly model relaxation
mechanisms.
The full Navier-Stokes equations are considered a viable alternative for many
systems as they include more physics than the Euler equations. However, the in-
creased complexity of the Navier-Stokes equations makes it much more dicult to
simulate than the Euler equations. Computational uid dynamics algorithms most
often used in aeroacoustics, can be implemented in two or three dimensions and
are nonlinear, but they do not include the eects of molecular relaxation, and are
computationally expensive [19, 20, 21, 22, 23]. Wochner et al. [24] use nonlinear
uid dynamics equations written in terms the total uid dynamic variables rather
than the acoustic variables to include the eects of shear viscosity, bulk viscos-
ity, thermal conductivity, and molecular relaxation to simulate nonlinear acoustic
phenomena.
Computational aeroacoustics studies noise generation via either turbulent uid
motion or aerodynamic forces interacting with surfaces. Although no complete
scientic theory of the generation of noise by aerodynamic ows has been estab-
lished, one of the most practical aeroacoustic analyses relies upon a so called wave
equation using Lighthills aeroacoustic analogy. Lighthill rearranged the Navier-
Stokes equations into an inhomogeneous wave equation, thereby making an analogy
between uid mechanics and acoustics [25, 26].
However, these approaches are also limited because of their reliance on tradi-
tional nite dierence algorithms, requiring considerable computational resources,
subject to numerical instability, and are all based on the continuum assumption.
The Knudsen number is used to distinguish the regimes where dierent gov-
erning equations of uid dynamics are applicable. Fig. (1.1) schematically draws
the limits of applicability of dierent continuum and particle methods. Continuum
methods are applicable in the small range where Kn < 0.05. For larger Kn, the
momentum and heat uxes in the Navier-Stokes equations cannot be written in
terms of macroscopic quantities and hence the set of equations is incomplete [27].
4
For even larger Kn, the uid can no longer be considered a continuum, thus the
Navier-Stokes equations do not hold. The Boltzmann equation is valid over the
entire Kn range.
Figure 1.1. Limits of applicability of various mathematical models to simulate uid
ow [1]
1.4 Particle methods
Particle methods that are based on the Boltzmann equation are valid for all Knud-
sen numbers. Therefore, particle methods are necessary for, but not limited to,
problems where the Knudsen number is greater than about 0.05.
Some of the dierent particle methods include the Molecular Dynamics (MD)
model [28, 29, 30], Monte Carlo methods [31], cellular automata [32, 33], discrete
velocity [34, 35], and the Lattice Boltzmann Method (LBM) [36, 37, 38, 39, 40].
The most fundamental approach is the Molecular Dynamics model. In this ap-
proach the particles move according to Newtons laws and particle interactions
are calculated by force potentials. This model can even account for quantum me-
chanical eects. Molecular dynamics is usually required to simulate dense gases or
liquids, therefore the expensive intermolecular force calculations make it unsuitable
for dilute gases.
The Monte Carlo approaches and the Boltzmann equation are derived from the
Liouville equation. The term Monte Carlo was adopted in the 1940s for meth-
ods involving statistical techniques, such as the use of random numbers to nd the
5
solutions to mathematical or physical problems [31]. The rst documented appli-
cation of using a Monte Carlo method is given by Georges-Louis Leclerc, Comte
de Buon in 1777 when he describes a Monte Carlo procedure that can be used
to evaluate from the throws of a needle onto a oor ruled with parallel straight
lines [41]. The probability that the needle would intersect a line is calculated by
randomly throwing the needle many times and calculating the ratio of the number
of throws intersecting a line to the total number of throws. This probability gives
an estimation of and is discussed further in Appendix A. Lord Kelvin used a
Monte Carlo method in 1901 to perform time integrals which appeared in his ki-
netic theory of gases [42]. However, it was only in 1940s that Monte Carlo methods
were developed enough to be used for solving engineering problems.
The cellular automata method considers the evolution of idealized particles
which move at unit speed from one node to another on a discretized temporal
grid [32, 33]. The discrete velocity method allows molecules to move in continuous
space, with velocities that are discrete, and non-uniformly distributed [34, 35]. The
Lattice Boltzmann method (LBM) was developed from cellular automata and dis-
crete velocity methods, and applies Boltzmann equation approximations to cellular
automata models. Particle motion using the LBM is restricted to a lattice, with
simple collision rules to conserve mass and momentum [36, 37, 38]. The LBM has
been successful in the study of one-dimensional sound propagation [39, 40]. Lim-
itations of some of these methods can be signicant due to the simplication and
discretization of physical or velocity space, and can include the inability to simu-
late heat transfer eects or the eect of temperature on the transport properties.
These limitations are discussed further by Long et al. [33].
The direct simulation Monte Carlo (DSMC) is a stochastic, particle-based
method rst introduced by G. A. Bird in 1963 [43] which is capable of simulat-
ing real gas eects for all values of Kn that traditional continuum models cannot
oer [1]. Particles in a DSMC simulation are not restricted to a grid and move
according to their velocities. While the collisions between particles are determined
statistically, they are required to satisfy mass, momentum, and energy conserva-
tion. DSMC oers computational exibility to study sound for all Kn in a wide
range of systems.
6
1.5 Sound for all Knudsen numbers
The Knudsen number is large for sound propagation in very dilute gases, in micro-
channels, or at high frequencies, and thus requires a particle method, or kinetic
theory solution. There have been numerous attempts to study sound propagation
for high Kn based on kinetic theory in simple monatomic gases [44, 45, 46, 47, 48,
49, 50, 51]. Most of these attempts use approximations that replace the Boltzmann
collision integral with concomitant losses in accuracy. All attempts report that the
speed and absorption of sound depend heavily on Kn which deviates signicantly
from traditional continuum theory at high Kn. Experimental results show close
agreement to kinetic theory results for Kn < 1, but for larger Kn, the results are
poor and inconsistent [4, 5, 52, 53, 54]. Detailed descriptions of continuum and
kinetic theory predictions for the absorption and dispersion of sound will be given
in Chapter (5).
Several authors have conducted theoretical studies of sound waves in rareed
polyatomic gases to include internal energy eects on the propagation of sound.
These studies are based either on classical or kinetic theory. Classical contributions
include Greenspan [6], Herzfeld and Rice [55], Herzfeld and Litovitz [3], and Kneser
[56]. The kinetic theory studies of interest are by Wang Chang and Uhlenbeck [57],
Monchick et al. [58], Mason and Monchick [59], Hanson and Morse [60], Hanson et
al. [61], McCormack and Creech [62], McCormack [63], Banankhah [64]. As in the
monatomic gas case, most of these attempts use approximations that replace the
Boltzmann collision integral. Results vary in degrees of success with comparison
to experimental studies in gases with internal energy [52, 53] and will be discussed
further in Chapter (6).
Because of the inconsistencies between theory and experiment at high Kn and
the limitations of continuum methods at low Kn, a computational method that
spans the whole Kn range is desirable. The best approach, one of the few available,
is to use the particle-based method DSMC to bypass continuum limitations while
also minimizing the numerical complexity of the model itself. DSMC has become
the de facto standard computational approach for high Kn due to its success at
accurately simulating a wide range of phenomena. In particular, sound propaga-
tion properties such as nonlinear phenomena and absorption are inherent in the
7
algorithm because of the particle nature of the algorithm. Therefore, the DSMC
method was chosen for study in its application to the simulation of the acoustic
wave propagation.
1.6 Direct Simulation Monte Carlo
1.6.1 Introduction
The direct simulation Monte Carlo (DSMC) method is a direct particle simula-
tion tool that describes the dynamics of a gas through direct physical modeling
of particle motions and collisions. DSMC is based on the kinetic theory of gas
dynamics modeled on the Boltzmann equation, where representative particles are
followed as they move and collide with other particles. Introductory [65, 66] and
detailed [1] descriptions of DSMC, as well as formal derivations [67] can be found
in the literature and will be discussed further in Chapter (3). Due to the particle
nature of the method, DSMC oers considerable exibility with regard to the type
of system available for modeling: rareed gas dynamics [1, 68, 69, 70], hypersonic
ows [1, 71, 72], and acoustics [65, 2, 73, 74]. DSMCs origins are based upon
the Boltzmann equation, but the applications of and the extensions to DSMC
method have now gone beyond the range of validity of this equation by simulating
chemical reactions [75, 76, 77] detonations [78, 79], and volcanic plumes and upper
atmospheric winds on Jupiters moon Io [80, 81].
DSMC is a particle method that describes the state of the gas at the micro-
scopic level, and is valid beyond the continuum assumption. In the case of one
dimensional acoustic wave propagation, the characteristic length is the acoustic
wavelength so the Knudsen number is directly proportional to the frequency of
oscillation. Despite the fact that DSMC is valid for all Kn, DSMC is most ef-
cient for high Kn ows. DSMC has traditionally been used in regimes where
continuum methods fail, but in fact compares well with Navier-Stokes calculations
and experimental results within the continuum regime [72]. Moreover, DSMC has
many advantages over traditional continuum approaches even for low Kn situa-
tions. Without modication, DSMC is capable of simulating all physical properties
of interest at the molecular level for sound propagation: absorption, dispersion,
8
nonlinearity, and molecular relaxation.
A parallel, object oriented DSMC solver was developed for this problem. The
code was written in C++ and is designed using the benets of an object oriented
approach [82, 83]. The parallelization is achieved by performing parallel ensemble
averaging, using the Message Passing Interface (MPI) library for inter-processor
communications [84, 85]. The reason for implementing a parallel model is to reduce
the high computation time, and make the problem more ecient. An investigation
of this is discussed in Chapter 4.
DSMC simulations in a simple gas at low amplitudes are described in Chapter
(5) and as a function of amplitude in Chapter (7) to investigate absorption and
dispersion as a function of Kn.
The DSMC program contains several types of energy models to treat molecules
in gas mixtures with internal energy. Internal energy is represented by rotational
and vibrational modes (electronic energy is ignored) and has been programmed
to simulate either classical or quantum behavior. Details on the internal energy
models used in DSMC will be given in Chapter (3). DSMC simulation results
implementing these internal energy models along with their eect on acoustics is
shown for a wide range of temperatures in Chapter (6) [2].
The exibility of the DSMC algorithm allows for modeling of sound in specic
gas mixtures including models for Earth, Mars and Saturns moon Titan [73, 86,
87, 88]. Little modication to the method is needed to change the molecular and
ambient atmospheric properties in order to simulate sound on the dierent planets.
This feature of DSMC makes it benecial for use in planetary acoustics where
atmospheric conditions are dependent on planet, time of year, altitude, etc. In
addition, the Kn is high in upper atmospheric conditions, thus requiring a particle
method solution. DSMC simulation results describing acoustic phenomenon on
Earth, Mars and Titan for all Kn is given in Chapter (8).
Chapter (9) is a summary of the work presented along with general conclusions
regarding this research. Potential applications and modications of the method
are also described.
Chapter 2
Kinetic Theory of Gases
2.1 Historical background and
Ludwig Boltzmann
Ludwig Boltzmann (1844-1906) was an Austrian physicist and philosopher and is
pictured in Fig. (2.1). In the late 1800s, Boltzmann claimed that matter was
made up of tiny particles, that is atoms and molecules. Much of the physics
establishment at the time did not share his belief, which was subject to much
debate. Boltzmann had many scientic opponents, and was even on bad personal
terms with some of his colleagues. One of the leaders of the anti-atom school,
Wilhelm Ostwald, often battled Boltzmann ercely on some of his theories. One
of the battles between the two was described by a colleague Sommerfeld [89].
The battle between Boltzmann and Ostwald resembled the battle of the bull
with the supple ghter. However, this time the bull was victorious...
Boltzmann suered from an alternation of depressed moods with elevated, ex-
pansive or irritable moods, and the often violent disputes about his scientic theory
were enough to depress Boltzmann. He began to feel that his lifes work was about
to collapse despite his attempts to defend his theories. He attempted suicide once
in 1900, only to succeed in taking his life in 1906 just before experimental work
veried his theories [90]. Perrins studies of colloidal suspensions (1908-1909) [91]
conrmed the values of Avogadros number and Boltzmanns constant, and Ein-
steins theory on Brownian motion (1905) [92] helped convince the world that the
10
Figure 2.1. Ludwig Boltzmann (1844-1906)
tiny particles really exist. Ostwald later told Sommerfeld that he had been con-
verted to a belief in atoms by Einsteins complete explanation of his theory on
Brownian motion [93].
Boltzmanns equation for entropy s = k
b
log W as it was published in his day,
is engraved on his tombstone, where k
b
is called the Boltzmann constant.
One of Boltzmanns main contributions is the invention of the eld of statistical
mechanics. In addition, Boltzmanns work in the foundations of kinetic theory
still remains widely popular as his theories including the distribution for molecular
speeds in a gas are applicable to many phenomena in gas dynamics, and will now
be discussed.
11
2.2 Velocity distribution function
Consider a simple gas in which all molecules are alike with mass m. Let dr denote
a small parcel of this gas surrounding the point in space r. This parcel is large
enough to contain many molecules but small enough so that ambient macroscopic
quantities such as pressure or temperature do not vary throughout the parcel. The
number of molecules within the volume dr averaged over a time dt is then given
by ndr where n is the number density. The number density, n, and mass density,
, are related by:
= nm. (2.1)
The distribution of velocities among a large number of molecules in the parcel
dr can be represented by the velocity distribution function, f, that describes the
state of the gas. As molecules pass in and out of the volume dr, the distribu-
tion function is a function of time. Therefore, the probable number of molecules
that have velocities in the neighborhood of u = (u
x
, u
y
, u
z
) and positions in the
neighborhood of r = (x, y, z), at the time t, is equal to:
f(u, r, t)dudr . (2.2)
Moreover, the whole number of molecules in the parcel dr is given by integrating
Eq (2.2) throughout the whole velocity space. Therefore,
n =
___
u
fdu. (2.3)
In this case of a simple gas with no internal energy, the independent variables
are the three components of particle velocity and position, and time, which dene
phase space. The particle velocity and position ranges from to in each
direction.
2.2.1 Macroscopic properties in a simple gas
Macroscopic ow quantities such as temperature and pressure can be calculated
in terms of averages over particle velocities. Therefore, we can use the velocity
12
distribution function to calculate all macroscopic quantities. For any quantity
Q(u) that is a function of molecular velocity u, the mean value of Q, denoted Q,
can be written as:
Q =
1
n
___
u
Qfdu. (2.4)
The lower-order moments of the distribution function all have names and simple
physical interpretations, where the k
th
moment is dened as:
___
u
mu
k
fdu, (2.5)
Because u is a vector, higher order moments can become tensors and require
tensor multiplication within the integrand. The tensor product, denoted , of
vector a = (a
1
, a
2
, a
3
) and b = (b
1
, b
2
, b
3
) is given as:
(a
1
, a
2
, a
3
) (b
1
, b
2
, b
3
) =
_
_
_
_
a
1
b
1
, a
1
b
2
, a
1
b
3
,
a
2
b
1
, a
2
b
2
, a
2
b
3
,
a
3
b
1
, a
3
b
2
, a
3
b
3
,
_
_
_
_
(2.6)
Using Eq. (2.3), the zeroth moment of the distribution function becomes:
nm = =
___
u
mfdu, (2.7)
giving the mass density of the gas .
The rst order moment of the distribution function denes the particle ux
density associated with the transport of mass. Using Eqs. (2.4) and (2.5), this is
written as:
u =
___
u
mufdu, (2.8)
where u = U is dened as the mean or stream velocity. The velocity of a molecule
relative to the stream velocity is call the thermal velocity, denoted by u

, is given
by:
u

= u U . (2.9)
The second moment case is of particular interest, dening the momentum ux
13
by the thermal motion of the gas. Since momentum is a vector quantity, the
resulting expression is a tensor with nine Cartesian components, called the pressure
tensor p and given by:
p = u

=
___
u
mu

fdu. (2.10)
In component form, the pressure tensor is given by:
p =
_
_
_
_
u
2
x
, u

x
u

y
, u

x
u

z
,
u

y
u

x
, u
2
y
, u

y
u

z
,
u

z
u

x
, u

z
u

y
, u
2
z
,
_
_
_
_
. (2.11)
Using tensor notation, the pressure tensor is written:
p
ij
= u

i
u

j
. (2.12)
The mean or hydrostatic pressure p is usually then dened as the average of
the three normal components of the pressure tensor, that is:
p =
1
3
(u
2
x
+ u
2
y
+ u
2
z
) =
1
3
u
2
. (2.13)
The viscous stress tensor is dened as the negative of the pressure tensor
p with the scalar pressure p subtracted from the normal components. It can be
written in tensor notation as:
=
ij
= (u

i
u

ij
p) , (2.14)
where
ij
is the Kronecker delta such that
ij
= 1 if i = j and
ij
= 0 if i ,= j.
The temperature T of a gas in uniform steady state is directly proportional to
the average kinetic energy associated with the translational motion of a molecule,
1
2
mu
2
. The proportionality is given by the relation:
1
2
mu
2
=
3
2
k
b
T , (2.15)
where k
b
= 1.3806504 10
23
[J/K] is the Boltzmann constant which is related
14
to the universal gas constant 1 = 8314.34 [J/(kmol K)] by k
b
= 1/^ where
^ = 6.02214179 10
26
[mol/kmol] is Avogadros number.
An immediate consequence of the denition of temperature is that the hydro-
static pressure p of a gas in equilibrium can be written in terms of the temperature
T resulting in the perfect gas law:
p = 1T/M = nk
b
T , (2.16)
where M = m/^ is the molecular weight of the molecule. T is an equilibrium
gas property, but Eq (2.16) will hold, even in nonequilibrium situations, for the
translational kinetic temperature T
tr
which is dened as:
3
2
k
b
T
tr
=
1
2
mu
2
. (2.17)
In addition, the translational temperature may also be dened for each velocity
component separately. That is, the translational temperature in the x direction
can be dened as:
k
b
T
trx
= mu
2
x
, (2.18)
which can therefore provide a measure of translational nonequilibrium when com-
pared to Eq. (2.17). Measuring the amount of nonequilibrium in a system is a
useful tool that is unavailable in traditional continuum methods.
For diatomic or polyatomic gases that possess internal energy, a temperature
T
int
can be dened for the internal modes associated with the rotational and vi-
brational energy of the molecule. Similarly to the translational temperature, the
internal temperature T
int
can be dened as:
1
2

1
M
T
int
= e
int
, (2.19)
where is the number of internal degrees of freedom and e
int
is the energy associ-
ated with the internal energy mode of interest. For a nonequilibrium gas, an overall
kinetic temperature T
ov
can be dened as a weighted average of the translational
and internal temperatures given by Eqs (2.17) and (2.19) given by:
T
ov
= (3T
tr
+ T
int
)/(3 + ) . (2.20)
15
Finally, the third moment of the velocity distribution function, the heat ux
vector q, is obtained. For a simple gas, the molecular energy is given by
1
2
mu
2
and the heat ux vector is then given by:
q =
1
2
u
2
u

=
___
u
1
2
mu
2
u

fdu. (2.21)
For a gas in equilibrium at constant volume, the amount of heat that is added
to the system when the temperature changes is called heat capacity at constant
volume, denoted C
v
, and is related to the universal gas constant by the equation:
C
v
=
1
2
, (2.22)
where is the number of degrees of freedom.
For a gas in equilibrium at constant pressure, the amount of heat that is added
to the system when the temperature changes is called heat capacity at constant
pressure, denoted C
p
, and is related to the heat capacity at constant volume by
the equation:
C
p
= 1+ C
v
. (2.23)
The specic heat at constant volume and pressure, c
v
and c
p
respectively, are
related to the molecular weight of the gas and are dened as:
c
v
=
C
v
M
(2.24)
and
c
p
=
C
p
M
. (2.25)
The ratio of specic heats is denoted by . Thus:
=
c
p
c
v
=
C
p
C
v
= 1 +
2

. (2.26)
where is the total number of degrees of freedom.
16
Eqs. (2.22) and (2.23) can now be written as:
C
v
=
1
1
, (2.27)
and
C
p
= 1
_
1 +
1
1
_
. (2.28)
In addition, the specic heats can also be written:
c
v
=
k
b

2m
. (2.29)
c
p
=
k
b
m
_
1 +

2
_
. (2.30)
2.3 Deriving the Boltzmann equation
Consider once more a small parcel of uid. At any particular instant in time, the
number of molecules in the phase space element dudr is given by Eq. (2.2). If the
location and shape of the uid element does not vary with time, the rate of change
of the number of molecules in the element is given by the expression:
f
t
dudr . (2.31)
There are three processes that can contribute to the change in the number of
molecules within the phase space element dudr. They are:
1. The movement of molecules by the molecular velocity u.
2. The convection of molecules as a result of an external force per unit mass,
F, which may be a function of r and t but not u.
3. The scattering of molecules as a result of intermolecular collisions.
When considering the rst two processes during the times between t and t +dt,
any molecule with velocity u will change to u+Fdt and its position r will change
to r+udt. At time t, there are f(u, r, t)dudr molecules in this set. Without taking
17
intermolecular collisions into account, after the time interval dt, these molecules
now occupy the uid parcel r +udt with velocities u +Fdt. Applying Eq. (2.2),
the number of such molecules in this set is given by:
f(u +Fdt, r +udt, t + dt)dudr . (2.32)
The dierence between these sets of molecules is the portion of Eq. (2.31) due to
scattering of molecules as the result of intermolecular collisions, which is denoted
(
col
f/t)dudr. Therefore,
f(u +Fdt, r +udt, t + dt)dudr f(u, r, t)dudr =

col
f
t
dudrdt . (2.33)
Dividing both sides by dudrdt gives:
f(u +Fdt, r +udt, t + dt) f(u, r, t)
dt
=

col
f
t
. (2.34)
Letting dt go to zero results in the equation to be written in terms of partial
derivatives:
f
t
+ u
x
f
x
+ u
y
f
y
+ u
z
f
z
+ F
x
f
u
x
+ F
y
f
u
y
+ F
z
f
u
z
=

col
f
t
(2.35)
Written in vector notation, this results in the Boltzmann equation:
f
t
+u
f
r
+F
f
u
=

col
f
t
. (2.36)
2.3.1 The calculation of the collision integral
In the evaluation of the collision integral, it is assumed that only binary collisions
between molecules are considered and collisions occupy a very small part of the
lifetime of a molecule. Consider at the moment, the collision of a molecule with
velocity u and a molecule with velocity u
1
. Their post collision velocities are
given by u

and u

1
respectively. Let a molecule with velocity u be chosen as a
test particle moving with speed u
r
= u u
1
amongst stationary molecules with
velocity u
1
. The number of molecules in the uid element dr that have velocities
of u
1
is given by f
1
du
1
. The volume swept out in physical space per unit time
18
that accounts for the cross section for this collision is given by u
r
d where is
the collision cross sectional area and d is the scattering angle. Therefore, the
number of collisions between molecules with velocity u
1
with the test particle per
unit time is given by:
f
1
u
r
ddu
1
, (2.37)
where f denotes the value of the velocity distribution function f at u. Similarly,
f
1
, f

and f

1
denote the values of f at u
1
, u

and u

1
respectively. Since the
number of molecules with velocity u in the phase space element dudr is fdudr,
the number of collisions between molecules with velocity u and u
1
resulting in
post collision velocities of u

and u

1
is:
ff
1
u
r
ddu
1
dudr . (2.38)
This expression represents the loss of molecules with velocity u due to collisions
with molecules with velocity u
1
. There are, however, inverse collisions that result
in molecules with post collision velocities of u. Similarly, the number of these such
encounters is:
f

1
u
r
ddu
1
dudr . (2.39)
The rate of increase of molecules that have velocity u in the phase space element
dudr as a result of direct and inverse collisions with molecules with velocity u
1
is
obtained by combining equations (2.38) and (2.39). This gives:
(f

1
ff
1
)u
r
ddu
1
dudr . (2.40)
The net gain of molecules per unit time that have velocity u is obtained by in-
tegrating this expression over the cross section for its collision with molecules of
velocity u
1
and then integrating over all velocity space. This results in:
dudr
___
u
1
4
_
0
(f

1
ff
1
)u
r
ddu
1
, (2.41)
19
which can be denoted by (
col
f/t)dudr. Hence, dividing by dudr results in the
collision integral:

col
f
t
=
___
u
1
4
_
0
(f

1
ff
1
)u
r
ddu
1
. (2.42)
Combining Eq. (2.36) and (2.42), the full expression for the Boltzmann equation
becomes:
f
t
+u
f
r
+F
f
u
=
___
u
1
4
_
0
(f

1
ff
1
)u
r
ddu
1
. (2.43)
2.4 Deriving conservation equations
from the Boltzmann equation
Despite the fact that the Boltzmann equation was derived using a microscopic
approach, it can also be shown that the continuum conservation equations are
moments of the Boltzmann equation. To see this, begin by multiplying both sides of
the Boltzmann equation by a quantity Q(u) that is a function of molecular velocity
and then integrating over velocity space. The multiplication of the Boltzmann
equation (2.36) gives:
Q
f
t
+ Qu
f
r
+ QF
f
u
= Q

col
f
t
. (2.44)
Q is a function of molecular velocity and not of distance, thus when integrating
both sides over velocity space we can write:
___
u

t
(Qf)du +
___
u
(Quf)du +
___
u
QF
f
u
du =
___
u
Q

col
f
t
du. (2.45)
20
The third integral term can be rewritten using the chain rule. In addition, by
assumption, F is independent of u, thus, Eq. (2.45) can be simplied as:
___
u

t
(Qf)du+
___
u
(Quf)du
___
u
F
Q
u
fdu =
___
u
Q

col
f
t
du. (2.46)
By using Eq. (2.4) this can be rewritten as:

t
(nQ) + (nQu) nF
Q
u
=
___
u
nQ

col
f
t
du. (2.47)
If the quantity Q is either the mass m, momentum mu or energy
1
2
mu
2
of
a molecule, then the collision integral

col
f
t
on the right hand side of Eq. (2.47)
is zero because the details of binary collisions say that mass, momentum, and
energy are conserved during collisions. The quantities m,mu and
1
2
mu
2
are called
summational invariants. It can be shown that these, or linear combinations of
these, are the only summational invariants [94].
The equation for the conservation of mass is obtained for substituting Q = m
into Eq. (2.47). This gives:

t
+ (U) = 0 . (2.48)
Next, by setting Q = mu the conservation of momentum equation becomes:
U
t
+ (U
2
) +p F = 0 . (2.49)
Eq. (2.49) can be rewritten in terms of the material derivative D/Dt =

t
+U
and the denitions of pressure and viscous stress tensor from Eqs. (2.13) and
(2.14). This gives:

DU
Dt
+p F = 0 . (2.50)
Finally, inserting
1
2
mu
2
in for Q results in the conservation of energy equation:

t
(
1
2
U
2
) + (
1
2
uu
2
) U F = 0 . (2.51)
21
The averages over molecular velocities in Eq. (2.51) may be converted to aver-
ages over the thermal velcoities. Introducing the heat ux vector dened by Eq.
(2.21), the hydrostatic pressure p given by Eq. (2.13) , and which is called the
dissipation function and is written in Cartesian coordinates as:
=
xx
U
x
x
+
yy
U
y
y
+
zz
U
z
z
+
xy
_
U
x
y
+
U
y
x
_
+
yz
_
U
y
z
+
U
z
y
_
+
zx
_
U
z
x
+
U
x
z
_
, (2.52)
the conservation of energy equation can be rewritten as:

De
Dt
= p U q +. (2.53)
Here, e is the intrinsic energy per unit mass.
Equations (2.48), (2.50), and (2.53) make up the Navier-Stokes equations. Note
that the original Navier-Stokes equation derived by Navier in 1822 and by Stokes
in 1845 is the momentum equation shown in Eq. (2.50) for a simple gas [93].
Due mainly to the popularity of computational uid dynamics, citations of the
Navier-Stokes equations have grown to include the mass and energy equations to
form the complete system of equations described above. Because the momentum
conservation equation (2.50) is a vector equation, the the Navier-Stokes equations
constitute ve equations. Together with the equation of state, in this case, the
perfect gas law given by Eq. (2.16), the set of equations is closed and can describe
the continuum behavior of a uid. If both and q are zero, then the momentum
equation reduces to Eulers equation given by:

DU
Dt
+p = 0 . (2.54)
Together with Eqs. (2.16) and (2.48), Eulers equation can describe continuum
behavior in the limit of Kn approaching zero.
22
2.5 Solutions to the Boltzmann equation
Because the Boltzmann equation is an integro-dierential equation, and because
particle velocities and positions must be known for all particles at all times, exact
solutions to the Boltzmann equation are complex and only exist in physically
simple situations. For problems in nontrivial gases or multidimensional ows, an
analytical solution to the Boltzmann equation is improbable, and even proves to
be dicult to obtain a computational solution using standard numerical methods
[95, 96]. Approximations and limited solutions to the Boltzmann equation are
discussed below.
2.5.1 Equilibrium properties
For a simple, monatomic gas that is subject to no external force eld and is sta-
tionary so that the velocity distribution is independent of r, Boltzmanns equation
(2.36) reduces to:
f
t
=
___
u
1
4
_
0
(f

1
ff
1
)u
r
ddu
1
. (2.55)
Boltzmanns H-theorem [15] shows that in order for this to be true, then:
___
u
1
4
_
0
(f

1
ff
1
)u
r
ddu
1
= 0 , (2.56)
or:
f

1
= ff
1
. (2.57)
This implies that the probable number of molecules in any element of velocity
space remains constant in time, that is, collisions between molecules do not aect
the velocity distribution function. Therefore, the velocity distribution function
here describes the equilibrium state of the gas.
It can be shown that solution to Eq. (2.57) requires f in its equilibrium state,
23
denoted f
0
, to take the form:
f
0
= n
_
m
2k
b
T
_
3/2
e
mu
2
/2k
b
T
, (2.58)
which is called the Maxwellian velocity distribution function. Eq. (2.58) can be
written in terms of the most probable molecular thermal speed u

m
=
_
2k
b
T/m as:
f
0
= n

3
m

(3/2)
e

2
m
u
2
, (2.59)
where
m
= 1/u

m
. The average thermal speed u

is related to the most probable


molecular thermal speed by u

= (2/

)u

m
.
Fig. (2.2) shows the Maxwellian distribution function for the monatomic gas
argon (molecular weight M = 39.9 kg/kmol) in equilibrium at 0

C with pressure
of 101,000 Pa. The Maxwellian distribution plotted in Fig. (2.2) gives the number
of molecules with speed u in a volume element around r.
Figure 2.2. Maxwellian velocity distribution function for argon at 0

24
As Kn goes to 0, the Maxwellian velocity distribution function provides a
solution to the Boltzmann equation. The range of validity of this solution is when
the gas is in equilibrium and can also be governed by the continuum equations of
Euler.
2.5.2 Linearized Boltzmann Equation
The theory based on work by Chapman [97] and Enksog [98] provides a perturba-
tion solution to the Boltzmann equation for a restricted set of problems [15]. In
this theory, a solution to the Boltzmann equation is represented by an asymptotic
power series expansion of succesive approximation. The expansion is carried out
assuming small deviations from the Maxwellian velocity distribution function f
0
dened by Eq. (2.58). The solution to the Boltzmann equation can be written in
the form of the power series:
f = f
0
+ f
1
+
2
f
2
+ . . . , (2.60)
where in this case is taken to be either the mean collision time or the Knudsen
number. Equation (2.60) is commonly rewritten as:
f = f
0
(1 +
1
+
2
+
3
+ . . .) . (2.61)
Upon assuming a solution of zeroth order, that is, f = f
0
, the Euler equations
are obtained. The Navier-Stokes equations are obtained by assuming a rst order
solution, f = f
0
(1 +
1
), and a second order approximation results in the Burnett
equations [99, 100].
Because the method of successive approximation implies that higher order
terms are dependent on lower order terms, the complexity of successive terms in-
creases rapidly with each iteration. In addition, the range of validity is restricted
at any order. For example, numerical instabilities and violations of the second law
of thermodynamics can occur with higher order solutions at high Kn [99]. How-
ever, to describe the gas dynamics in the Earths atmosphere near the ground, a
rst order solution is a suciently good approximation for most purposes [15].
25
2.5.2.1 1st order solution
Substitution of f = f
0
(1 + ) into the Boltzmann equation with no force eld
(2.36) yields:

t
+u

r
=
___
u
1
4
_
0
f
0
f
0
1
( +
1

1
)u
r
ddu
1
, (2.62)
which can be shown [15] to have a solution of the form:
=
1

m
n
A ln T
2
n
B : U (2.63)
where B : U denotes the scalar product of B and U. This can also be written
in tensor notation B
i,j
U
i
x
j
. A and B are functions of T and U and are usually
written as an innite series of Sonine polynomials S
i
[15]. That is:
A =

i=1
a
i
S
i
U , (2.64)
and
B =

i=1
b
i
S
i1
_
|
i
|
j

|
2
3

ij
_
, (2.65)
where | =
_
m
2k
b
T
_
1/2
U
Sonine polynomials arise in the study of Bessel functions and are solutions to
the dierential equation [101, 102]:
xy

+ (m + 1 x)y

+ (n m)y = 0 , (2.66)
where m ,= 0 and n is a real number. Sonine polynomials were rst used in
the kinetic theory of gases by Burnett [103]. The ease of certain orthogonality
properties and fast convergence makes Sonine polynomials ideal for use in this
context [103]. The results by Burnett are essentially identical to the earlier work
by Chapman [97] and Enskog [98] without the elaborate expansions that were
necessary in Chapman and Enskogs theory.
26
2.5.2.2 Transport coecients
The innite sums A and B from Eqs. (2.64) and (2.65) can be rewritten since all
terms vanish except for i = 1 using properties of Sonine polynomials:
A = a
1
_
5
2
|
2
_
U , (2.67)
and
B = b
1
_
|
i
|
j

|
2
3

ij
_
. (2.68)
The coecient of thermal conductivity and the coecient of viscosity can be
written in terms of A and B given above. It can be shown that rst approximation
to the coecient of thermal conductivity for a simple monatomic gas is:

(1)
=
5c
v
k
b
T
3
a
1
. (2.69)
Similarly, the rst approximation to the coecient of viscosity is written:

(1)
= b
1
k
b
T . (2.70)
Evaluating a
1
and b
1
from the Sonine polynomial expansion for a hard sphere
molecular model, Eq. (2.69) can be written in terms of the coecient of viscosity
by:

(1)
=
5
2
c
v

(1)
. (2.71)
For a gas with molecular diameter d, the rst approximation to the coecient of
viscosity is then:

(1)
=
5
16d
2
_
k
b
mT

_
1/2
. (2.72)
Later approximations can be determined by evaluating the Sonine polynomials
further. To the fourth degree of approximation:

(4)
= 1.02513
(1)
, (2.73)
and

(4)
= 1.01600
(1)
. (2.74)
27
Under the hard sphere molecular model approximation, the viscous stress tensor
can now be written:

i,j
=
_
U
i
x
j
+
U
j
x
i
_
+
2

i,j
U
k
x
k
, (2.75)
where
2
is the second coecient of viscosity, which is related by the bulk viscosity
dened as
b
=
2
3
+
2
.
The heat ux vector also now becomes:
q = T . (2.76)
In the case of a dilute monatomic gas with no internal energy, the bulk viscosity
is zero, so the viscous stress tensor can be rewritten as:

i,j
=
__
U
i
x
j
+
U
j
x
i
_

2
3

i,j
U
k
x
k
_
. (2.77)
Inserting Eqs. (2.67) and (2.68) into Eq. (2.63), the Chapman Enskog solution
for can be written:
=
a
1

m
n
_
5
2
|
2
_
U ln T
2b
1
n
_
|
i
|
j

|
2
3

ij
_
U
i
x
j
(2.78)
Written in terms of Eqs. (2.69) and (2.70), Eq. (2.78) can be rewritten as:
=
1

m
n
3
(1)
5c
v
k
b
T
_
|
2

5
2
_
U ln T
2
n

(1)
k
b
T
_
|
i
|
j

|
2
3

ij
_
U
i
x
j
. (2.79)
Using the perfect gas law and Eq. (2.71) along with f = f
0
(1 +), Eq. (2.79) can
be simplied to:
f = f
0
_
1
_
1

m
3
(1)
2p
_
|
2

5
2
_
Uln T +
2
(1)
p
_
|
i
|
j

|
2
3

ij
_
U
i
x
j
__
. (2.80)
Eq. (2.80) is referred to as the rst order Chapman-Enskog solution to the Boltz-
mann equation for a perfect monatomic gas.
Many assumptions and approximations have gone into the theory for the rst
order solution which is equivalent to the Navier-Stokes equations. To summarize,
28
these approximations are listed below:
1. Small perturbation from equilibrium
2. Monatomic gas
3. Perfect gas
4. Hard sphere gas
5. Continuum approximation
6. Bulk viscosity is zero
The following excerpt from Dorrance [104] appropriately describes the limi-
tations of using the continuum approach: The Navier-Stokes equations and the
remaining equations can be used to describe many situations with remarkable suc-
cess over a wide range of gas pressures and temperatures for the ow of mixtures
of polyatomic gases. We must not expect too much of the theory, however. There
is no a priori reason to expect that the theory can be applied to ionized gas ows
or the ow of liquids with any degree of success, nor is uncritical application of
the low-density or nearly free molecule ow of gases warranted.
2.5.3 Bhatnagar, Gross, and Krook (BGK) equation
Some other techniques have evolved which provide approximations to the Boltz-
mann equation as a method of nding a solution. One such technique is simplifying
the collision integral, as was done by Bhatnagar, Gross, and Krook [105]. In this
theory, it is assumed that collisions between molecules cause the velocity distribu-
tion function f to approach a Maxwellian value. This collision theory assumes, as
its basic approximation, that during time dt, a fraction dt/
c
of molecules undergo
collisions, which alter their velocity distribution function from f to f
0
where
c
is the time between collisions. In this case, the collision term of the Boltzmann
equation given by Eq. (2.42) can be approximated by the sum of two parts: the
loss term and the gain term. In the absence of an external force eld, this is written
as:
f
t
+u
f
r
= ((f) fL(f) . (2.81)
29
The simplest model for the gain and loss terms is:
((f) = /n, (2.82)
and
L(f) = /nf
0
, (2.83)
where / =
m
/u

m
for a hard sphere monatomic gas in equilibrium. u

m
is the
mean thermal speed of the gas. In this case, the Boltzmann equation reduces to:
f
t
+u
f
r
= /n(f
0
f) =
f
0
f

c
. (2.84)
Eq. (2.84), often referred to as the BGK equation, follows theory that says the
collision term should vanish when in equilibrium. This approximation to the Boltz-
mann equation is much simpler to model and solutions are more readily available
[96, 95]. However, its range of validity is much smaller than that of the full Boltz-
mann and Navier-Stokes equations and should only be used in near equilibrium
conditions [105].
2.5.4 Numerical solutions to the Boltzmann equation
When analytical solutions to the Boltzmann equation are not available or too
complex, numerical solutions are considered. Standard numerical method solutions
only exist for simplied geometries or approximations to the Boltzmann collision
integral by use of the BGK equation [95, 96].
As an illustration of the diculty involved in the computational problem for
solving the Boltzmann or BGK equation, consider a simple gas in which the Boltz-
mann equation is written in terms of phase space. This means that the computa-
tional problem is seven dimensional (space, velocity space and time) and discretiz-
ing phase space quickly becomes intractable. If one needs a 100 by 100 by 100 grid
in space and if even a small 20 by 20 by 20 grid could be used in velocity space,
this would require 8 billion grid points. Even if each grid point requires only 500
bytes of memory, this would mean 4 terabytes of memory for the whole system. In
addition, the computer operations would be on the order of N
7
and that computa-
30
tional cost is prohibitive. In addition, in order to model multiple molecular species
and chemical reactions, one would need separate Boltzmann equations for each
species. This curse of dimensionality is the reason why Monte Carlo methods
are the best approach for problems involving many dimensions [106].
The Direct Simulation Monte Carlo (DSMC) method has been shown to be
equivalent to a numerical solution of the Boltzmann equation [67]. DSMC was
derived through physical reasoning just as the Boltzmann equation was derived.
Because of its particle nature, DSMC can be applied to complex situations that
would not be accessible with standard numerical models. Chapter (3) presents the
DSMC method in detail.
Chapter 3
Implementation of DSMC for
Acoustics Simulations
3.1 Algorithm
The DSMC approach used for the simulation of acoustic wave propagation consists
of essentially four subroutines which are uncoupled during each time step t:
1. Move particles according to the time step and their velocity.
2. Apply boundary conditions, sort and index particles into cells.
3. Collide particles locally and allow for redistribution of internal energy if
simulating diatomic or polyatomic gases.
4. Sample macroscopic properties of interest in each cell.
Fig. (3.1) is a owchart of a typical DSMC simulation. Each step is called
systematically during every time step and each run is completed independently
with a new set of random numbers generated for each additional run. The number
of samples and the number of time steps between samples is determined the the
user dened values of NST and NIS respectively. So there are NST NIS total number
of time steps. The individual processes within the algorithm are described below.
32
Figure 3.1. A owchart of the DSMC algorithm. There are NST total samples with NIS
time steps in between samples. J and I increment NST and NIS respectively.
33
3.1.1 Initialization
Similar to a conventional numerical uid dynamics scheme, a computational do-
main in physical space is dened for DSMC simulations. An overlaying grid of cells
is used for selecting collision pairs and also for sampling macroscopic properties.
The rst step in the DSMC routine is to specify the initial state of the gas by
the initialization of cells and particles. Each cell in the computational domain is
initialized to contain a fraction of the total number of particles in the simulation.
In the cases considered here, every cell initially contains the same number of par-
ticles. The particles are initialized by specifying their molecular weight, diameter,
number of internal degrees of freedom, internal energy, location in space and veloc-
ity. The hard sphere molecular diameter is usually calculated by using Eq. (2.72)
and experimental values of viscosity. In the cases considered here all particles
within the domain are initialized to have random velocities sampled from the equi-
librium Maxwellian velocity distribution function dened by Eq.(2.58) for a given
temperature. The procedure for randomly sampling a Maxwellian distribution is
shown in Appendix (B). Each particle is also assigned internal energy depending
on how many internal degrees of freedom are active for the particular gas species
simulated. In most cases considered in this work, the gas was initialized to be in
thermal equilibrium so that the internal energy was dependent on the equilibrium
gas temperature.
3.1.2 Move particles and Boundary Conditions
Each particle is moved a distance ut every time step, where u is the velocity of
the particle and t is the global time step. Appropriate conditions are applied
for each molecule if it encounters a boundary. Some boundary conditions that
can be included are: diuse boundaries, specular boundaries, periodic boundaries,
inow/outow boundaries, and moving boundaries, in particular a piston boundary
condition.
Specular boundaries are mirror-like where the normal components of particle
velocity are reversed, the tangential velocity is retained and the energy is preserved.
The specularly reecting wall is analogous to an inviscid, slip boundary condition.
In a diuse reection, the particle that bounces o a wall is scattered into the
34
domain uniformly in all directions. Using diuse boundaries is equivalent to a
nonslip boundary, which is useful for simulating wall friction, and heat transfer
from wall to uid, for example. For periodic boundaries, all the particles that
exit the domain on one side are inserted back into the domain from the opposite
boundary. Inow/outow boundaries usually are based on Riemann invariants to
avoid macroscopic reections from the boundaries. Moving boundaries are capable
of simulating Couette ow and piston like systems [74].
In all the simulations reported in this work, acoustic waves were simulated in
a one dimensional simulation domain by creating a piston-like boundary condition
at one end of the domain. The piston was simulated as a rigid wall where particle
collisions with the piston face would result in sinusoidally oscillating velocity com-
ponents. The far end of the simulation domain was terminated with a specular
wall boundary condition. Even though the simulation domain is a one dimensional
array of cells, each particle still has three components of velocity and position.
Once the frequency of oscillation of the piston face is specied, the size of the
cells is determined so that there are sucient number of points per wavelength.
Then, given the time step size, the frequency and amplitude of oscillation, the pis-
ton boundary condition reects sinusoidally oscillating velocity components when
particles hit the boundary.
This simulation setup resembles the experimental setup seen in the literature
[5, 107] with a few modications. First, the simulated acoustic source is a rough
approximation of a transducer face in that the physical motion of the wall is
not modeled. The second dierence is that all macroscopic quantities (including
temperatures associated with internal energy of the molecules) can be computed
everywhere in the simulation domain whereas in experimental measurements, only
one macroscopic quantity is measured and only at the few locations where out-
put receivers are located within the domain. Therefore, it is not necessary with
DSMC to vary the domain length to obtain attenuation and velocity values, as in
the interferometry techniques used in experimental measurements. In addition, it
has been shown that the acoustic eld is altered by the introduction of receiving
transducers, and is dependent on the size of the transducer face [108]. In contrast
to the receivers used in experiments, the simulation setup is nonintrusive, which
becomes important in high Kn situations.
35
3.1.3 Sort particles
After the boundary conditions are taken into account, the particles are then sorted
to determine which cell they are in, which is a prerequisite step for modeling
collisions and sampling. It is important in this step to create a linked list that
includes how many particles are in each cell, as well as the cell location of each
particle, which is quite easy using C++ pointers. This list is updated every time
step. For a one-dimensional system, the equation for nding the cell number in
the direction of particle i with position x
i
and cell size x is given by:
cell number =
x
i
x
. (3.1)
This step is quite trivial for one-dimensional structured grids, but can become
expensive and dicult for multi-dimensional unstructured grids.
3.1.4 Collision
The next step, and the most unique to DSMC, is the collision routine. The collision
process is stochastic and sets DSMC apart from other particle methods. This
process assumes that intermolecular collisions are binary collisions involving just
two particles. The details on binary collisions are given below.
3.1.4.1 Binary collisions
Consider two particles that have pre-collision velocities of u
1
and u
2
respectively.
The post collision velocities, denoted u

1
and u

2
must reect the conservation of
linear momentum and energy. This requires:
m
1
u
1
+ m
2
u
2
= m
1
u

1
+ m
2
u

2
= (m
1
+ m
2
)u
m
, (3.2)
and
m
1
u
2
1
+ m
2
u
2
2
= m
1
u
2
1
+ m
2
u
2
2
. (3.3)
Here, m
1
and m
2
are the masses of the two molecules and u
m
is the velocity of
the center of mass of the molecule pair, which remains unaected by the collision.
36
The relative velocity pre and post collision can be dened as:
u
r
= u
1
u
2
, (3.4)
and
u

r
= u

1
u

2
. (3.5)
Combining Eqs. (3.2) and (3.4) gives:
u
1
= u
m
+
m
2
m
1
+ m
2
u
r
, (3.6)
and
u
2
= u
m

m
1
m
1
+ m
2
u
r
. (3.7)
Similarly, the post collision velocities can be determined from Eqs. (3.2) and
(3.5) as:
u

1
= u
m
+
m
2
m
1
+ m
2
u

r
, (3.8)
and
u

2
= u
m

m
1
m
1
+ m
2
u

r
. (3.9)
Eqs (3.6) - (3.9) show that:
m
1
u
2
1
+ m
2
u
2
2
= (m
1
+ m
2
)u
2
m
+ m
r
u
2
r
, (3.10)
and
m
1
u
2
1
+ m
2
u
2
2
= (m
1
+ m
2
)u
2
m
+ m
r
u
2
r
, (3.11)
where
m
r
=
m
1
m
2
m
1
+ m
2
, (3.12)
is the reduced mass. A comparison of Eqs. (3.10), (3.11), and (3.3) shows that the
magnitude of the relative velocity is unchanged by the collision in the absence of
internal energy exchange. That is:
u
r
= u

r
. (3.13)
37
Since both u
m
and u
r
can be calculated from the pre-collision velocities, the de-
termination of the post-collision velocities reduces to the calculation in the change
of direction of the relative velocity vector and Eqs. (3.8) and (3.9).
3.1.4.2 DSMC collision routine
Two particles that are in the same cell are chosen at random. The probability
of these two particles colliding over the time interval t is proportional to the
product of their relative speed u
r
and the collision cross section , i.e.
u
r
(u
r
)
max
, (3.14)
where (u
r
)
max
is the maximum collision cross-section and relative velocity product
encountered during the collision process. This value is stored in each cell and is
updated when a larger value is encountered.
If this probability is larger than a random number having a rectangular dis-
tribution in [0, 1], then the two particles collide, but if the probability is smaller
then two new particles are chosen at random until the condition is satised. Post
collision velocities are then computed assuming random scattering, but satisfying
conservation of energy and momentum given by Eqs. (3.8) and (3.9). For a gas
with internal energy, the energy exchange between the molecules occurs during in-
termolecular collisions and is also reected in the post collision velocity. Detailed
information about gases with internal energy will be shown in Section (3.2).
This collision process continues until the appropriate number of collisions has
reached the value:
1
2
N
c
N
c
F
num
(u
r
)
max
t
(x)
3
. (3.15)
where F
num
is the ratio of real to simulated particles and N
c
are the number of
particles in a cell. Eq. (3.15) is based on Eq. (2.38) which was derived for the
velocity distribution function. A ow diagram summarizing the collision routine
is shown in Fig. (3.2).
38
Figure 3.2. A owchart of the collision routine in the DSMC algorithm. sigVmax is
calculated by looping through all the particles in the cell, the number of collisions to
be performed, numColl, is given by Eq. (3.15) and the probability of collision, prob, is
given by Eq. (3.14). Details of the internal energy exchange routine will be given in Sec.
(3.2).
39
The collision model that is used in this research is the hard sphere model where
particles have xed molecular diameters. Dierent models are available for a more
exact representation of intermolecular collisions. They are: variable hard sphere,
soft spheres, and variable soft spheres [1]. These models represent the molecular
characteristics of gases with the added complexity of the eects of temperature
and/or attractive potential on molecular interactions. These molecular models
become important when there are large changes in temperatures to accurately
model viscosity. The changes in temperature in this work are relatively small and
the hard sphere molecular model is sucient for these purposes.
3.1.5 Sampling
Macroscopic properties are sampled every few time steps. The sampling is per-
formed by averaging over the number of molecules in each cell and their velocities.
Macroscopic quantities are computed using the particle velocity components in an
analogous way that macroscopic quantities are dened from the velocity distribu-
tion function in Sec. (2.2.1). As an example, if there are N
c
particles in a cell, the
density , stream velocity U, and translational temperature T
tr
for each cell are
given by the following equations:
= F
num
N
c
m
(x)
3
, (3.16)
U =
1
N
c
Nc

i=1
u
i
, (3.17)
_
_
_
_
T
trx
T
try
T
trz
_
_
_
_
=
m
N
c
k
b
Nc

i=1
(u

i
)
2
. (3.18)
Eq. (3.18) can be rewritten in terms of molecular velocity u and not the thermal
velocity u

. This gives:
_
_
_
_
T
trx
T
try
T
trz
_
_
_
_
=
m
N
c
k
b
_
Nc

i=1
u
2
i
U
2
_
40
=
m
N
c
k
b
_
Nc

i=1
u
2
i

_
1
N
c
Nc

i=1
u
i
_
2
_
(3.19)
T
tr
can be written as an average as its components so that T
tr
=
1
3
(T
trx
+T
try
+
T
trz
. If the system is in equilibrium then T
tr
= T
trx
= T
try
= T
trz
For a gas with internal energy, the internal temperature can be computed if
the number of internal degrees of freedom, , and internal energy e
int
are recorded.
The internal temperature T
int
is given by the equation:
T
int
=
2
k
b
Nc

i=1
e
int,i
Nc

i=1

i
. (3.20)
The overall temperature is then calculated using Eq. (2.20).
In a gas mixture, macroscopic quantities can also be computed for each indi-
vidual molecular species. The quantities given by Eqs. (3.16 - 3.20) are calculated
at the geometric center of each cell and are dened to be the average in that cell.
This sampling is not performed every time step.
3.2 Diatomic and polyatomic gases
The current DSMC program contains several types of energy models to treat
molecules in gas mixtures with internal energy. Internal energy is represented
by rotational and vibrational modes (electronic energy is ignored) and has been
programmed to simulate classical or quantum behavior. Each case uses a phe-
nomenological approach developed by Borgnakke and Larsen [109] which treats
only a fraction of intermolecular collisions as inelastic. This fraction is the re-
ciprocal of the relaxation collision number, Z, and is specied independently for
rotational and vibrational degrees of freedom and is dependent on the molecular
species of the colliding pair. If a collision is regarded as inelastic, the total en-
ergy e = e
int
+ e
tr
is reassigned between the translational and internal modes by
sampling from known equilibrium distributions during the collision but the total
energy is conserved [1, 109].
41
In the case of molecules having vibrational internal degrees of freedom, post
collision energy is assigned through either a classical procedure that assigns a con-
tinuously distributed vibrational energy to each molecule, or through a quantum
approach that assigns a discrete vibrational level to each molecule. The vibrational
energy exchange is treated independently from the energy exchange in rotational
modes. With DSMC each discrete vibrational energy level of a molecule can be
modeled as a separate molecular species. The discrete exchange model used is
based primarily on those developed by Anderson et al. [110] and Bergemann et
al. [111] and uses characteristic temperatures (or energy levels) and collision en-
ergies to determine the populations of excited states for a given molecular species.
This phenomenological model satises detailed balance and produces statistically
accurate macroscopic behavior in addition to being very computationally ecient
[1].
It is important to note that the energy exchanges are completed for each ac-
cepted molecule independent of its colliding partner. In general, the redistribution
of energy rst considers any available vibrational modes, and then rotation, then
the remainder is left as translation energy. Internal energy exchange is also reected
in the post collision relative velocity of the collision pair. Once the internal energy
exchange has occurred, the classical or quantum energy model will determine the
relative velocity and post collision velocity components can be determined from
Eqs. (3.8) and (3.9). A owchart describing the implementation of the internal
energy exchange during the collision routine is shown in Fig. (3.3).
42
Figure 3.3. A owchart of the internal energy exchange routine in the DSMC algorithm.
43
Given the temperatures under consideration in this study, it is assumed that
the rotational internal energy of the molecules is continuous, and thus is always
modeled using a classical procedure. The classical model is valid for rotation be-
cause quantum energy levels for rotation are closely spaced for the molecules of
interest in this study. The quantum energy levels for vibration are spaced more
widely apart. At low temperatures, most molecules will be in the ground vibra-
tional state, and a quantum description of vibration is required. At temperatures
on the order of the characteristic temperature of the gas, more molecules will be
excited to higher order states where a classical description of the vibrational en-
ergy may be sucient. Details about the eect of these internal energy models on
acoustics is described for a wide range of temperatures in Chapter (6).
3.2.1 DSMC tests for gases with internal energy
The results of a simple experiment of molecular relaxation in a nitrogen-like sys-
tem are shown in Figs. (3.4) and (3.5). The gas is initially perturbed from
a Maxwellian distribution by imposing a zero internal temperature. Fig. (3.4)
shows the fully classical rotation-vibration model model by plotting the internal
and translational temperatures as a function of number of collisions per molecule.
Fig. (3.4) shows the classical rotation-quantum vibration model by plotting the
rotational and quantum vibrational temperatures as a function of number of col-
lisions per molecule. Both simulations were completed under the same conditions
in a single cell with a cell length of 1/2 the mean free path, an initial transla-
tional temperature of 4000 K, and an average of 1 out of every 5 collisions treated
as inelastic. Relaxation is the mechanism at which the system exchanges energy
between translational and internal modes in order to reach a state of thermal equi-
librium. The number of collisions required to bring the system approximately to
equilibrium, also known as the relaxation collision number (Z), was chosen to be
5 for this case.
Both energy exchange models are compared to simple theoretical treatments
(exponential model rates) derived by Bird [1]. Fig. (3.4) shows excellent agreement
between the exponential model and the classical rate of energy exchange. In Fig.
(3.5), if no quantum behavior were present, the relaxation rate would be expected
44
Figure 3.4. DSMC simulation of nitrogen molecules undergoing classical relaxation at
4000 K. Birds exponential model[1] given by solid lines and DSMC results in dashed
lines [2].
to match the model rate. The coupled classical rotation/quantum vibration model
deviates from this idealized exponential rate and in fact gives a more realistic rate
of exchange than that of the fully classical approach. After a sucient number of
collisions, the distribution of velocities of the gas returns to the local Maxwellian
distribution in both cases.
3.3 Mixtures implementation
The current DSMC algorithm is written to simulate sound in a gas mixture. Eqs.
(3.16 - 3.20) can be computed for each molecular species or for the overall quantity
in each cell. For example, the overall number density, n, is related to the number
densities of all the individual species number densities. For a gas mixture of s
45
Figure 3.5. DSMC simulation of nitrogen molecules undergoing relaxation with a cou-
pled discrete vibration / classical rotation model at 4000 K. Birds exponential model[1]
given by solid lines and DSMC results in dashed lines [2].
separate species, the number density, is written:
n =
s

i=1
n
i
. (3.21)
In addition, the relaxation collision numbers can be specied independently for
each colliding pair. This gives a two dimensional array of collision numbers for
each combination of species pairs.
3.4 Assumptions and Error within DSMC
Some key DSMC assumptions such as molecular chaos are shared with the Boltz-
mann equation where particles are randomly distributed without correlation be-
tween velocity and position. In addition, the dilute gas assumption implies that
only binary collisions between molecules are possible and that collisions occupy
46
very little time in the life of a particle. This implies that in DSMC the movement
and collisions of particles is decoupled for a small time step. This assumption is
correct only when molecular chaos is assumed and the time step is less than the
mean collision time of the particles. It has been shown that the error in the trans-
port coecients dened in Section (2.5.2.2) is proportional to the square of the
time step, which for accurate simulations has to be much smaller than the mean
collision time [112, 113].
In addition, nite cell sizes are used to discretize the continuous physical space.
Alexander et al. have shown that the transport coecients deviate from kinetic
theory values proportional to the square of the cell size [114, 115]. Cell sizes on the
order of a mean free path can produce an error on the order of 10%. Therefore,
cell sizes are generally restricted to x 0.5
m
so the error is less than 2.5%.
Another computational assumption is that each simulated particle is assumed
to represent a number of real molecules, given by the ratio F
num
. If F
num
is large,
then there is a large amount of statistical scatter that can be many orders of
magnitude greater than the scatter that is present in the real gas. The statistical
scatter is reduced by ensemble averaging a large number of independent (randomly
seeded) realizations of an unsteady ow. The scatter or uctuations resulting in
simulations with F
num
= 1 is physically real and should not be considered an
error. The scatter is inversely proportional to the square root of the total number
of simulated particles and the computation time is directly proportional to the
sample size. So, while in principle it is possible to achieve any desired level of
accuracy by increasing the sample size, the convergence is slow and computation
time large by increasing the sample size. This unavoidable scatter has been a
major drawback for the DSMC method. The reduction in the cost of computing
means that it is easier to obtain accurate results, but simulating small deviations
from equilibrium can still be relatively computationally expensive. Modications
to DSMC in an attempt to address these drawbacks will be further discussed in
Chapter (4).
As an example, a system of argon gas at 0

C was initialized in equilibrium


in order to test the numerical convergence for 1, 10, 100, and 1000 independent
ensembles. Figs. (3.6) - (3.9) show the initial velocity components of this test over
the computational domain.
47
Figure 3.6. Equilibrium initialization of argon at 0

C after 1 ensemble
Figure 3.7. Equilibrium initialization of argon at 0

C after 10 ensembles
48
Figure 3.8. Equilibrium initialization of argon at 0

C after 100 ensembles


Figure 3.9. Equilibrium initialization of argon at 0

C after 1000 ensembles


49
The magnitude of the statistical error was also found to depend on the gas
temperature and gas type: lower temperatures and/or larger molecular masses
result in smaller statistical errors [116].
DSMC simulations become more exact as the time step and the cell size ap-
proach zero. Wagner [67] has proved that the DSMC method converges to the
solution of the Boltzmann equation in the limit of the time step and the cell size
equal to zero. Since the motion of the particles is independent of the cell struc-
ture, the disturbances can propagate at the sound or shock wave speed even when
the ratio of cell size to the time step is relatively very small. Therefore, DSMC
is not limited by a stability criterion such as the Courant-Fredrichs-Lewy (CFL)
condition.
Chapter 4
Computing Issues
4.1 Object-Oriented Programming Approach
The DSMC code used in this study was initially written by Prof. Lyle N. Long
[117] in C++ using an object-oriented programming (OOP) approach. Many mod-
ications have been made to expand the utility of this program, but has been
done maintaining the object-oriented language. Such an approach is implemented
through the use of object oriented properties of inheritance, polymorphism, and
encapsulation which enables the design of programs that are signicantly easier
to read, maintain, modularize, etc., than the typical top-down design approach
used in many C or FORTRAN programs. An object-oriented approach is most
suitable for a particle method such as DSMC because the particles and cells are
physical objects that have a dened set of attributes and functions. The program
is also more ecient by using C++ characteristics such as memory handling, class
structure and pointers [83, 82].
A class structure breaks the problem down into physical objects instead of
abstract quantities such as density and pressure. The cell class represents the
computational domain. Each cell object has its own dimensions, contains particle
objects and the cell class governs several methods (i.e. functions) that would occur
in a cell, such as the collision and sampling methods. A cell object only needs to
keep track of which particles belong to it and can then pass on information about
how the particles collide, and output the macroscopic properties of interest. The
details of such functions are hidden from the user.
51
Each particle object is responsible for tracking their own position and velocity.
This data is known only to the objects themselves as it is private data in the
particle class. In general a class can send private data to other objects that might
require that data, but outside objects are not able to directly alter this data.
The particle class also contains methods that are specic to the particle object.
For example, the move function is dened within the particle class, and is written
to determine how the particle moves based on its velocities. To the user, the details
of how the particle moves is not accessible, but the user can simply tell particle i to
move during the time step dt by writing part[i].move(dt); In this way, specic
aspects of the code can be modied without disrupting other parts of the code or
changing the structure of other objects. This encourages code reuse, since users
can use the code without worrying about the details or inner workings.
In a gas mixture, a species class is constructed to contain data pertinent to each
molecular species within the gas. For example, the molecular weight, diameter,
characteristic temperature, degrees of freedom, as well as the relaxation collision
numbers needed for internal energy exchange are all data members of the species
class.
Using these objects, the main time stepping routine of the direct simulation
Monte Carlo method ows in a logical sequence and is easy to understand. For
example, the time stepping algorithm is given below where NST is the number of
samples, NIS is the number of time steps per sample, NM is the number of molecules
in the system, DT is the time step size, NC is the number of cells, and the boundary
conditions are determined by xMinBC and xMaxBC for example.
52
do{
for (J = 0; J < NST; J++) { // LOOP OVER SAMPLES
for (I = 0; I < NIS; I++) { // LOOP OVER TIME STEPS
// PER SAMPLE
// MOVE ACCORDING TO TIME STEP SIZE AND APPLY BOUNDARIES
for (M = 0; M < NM; M++) {
atoms[M].move( DT );
atoms[M].xFarField( xMinBC, xMaxBC, DT );
atoms[M].yFarField( yMinBC, yMaxBC, DT );
atoms[M].zFarField( zMinBC, zMaxBC, DT );
}
// SORT PARTICLES DEPENDING ON WHAT CELL THEYRE IN
for (N = 0; N < NC; N++) { cells[N].zeroOutAtoms();}
// ADD PARTICLE M TO CELL K
for (M = 0; M < NM; M++) {
K = atoms[M].findCell( CH );
cells[K].addAtom ( &atoms[M] );
}
// COLLIDE
for (N = 0; N < NC; N++) {
cells[N].setMaxSigVel();
cells[N].collide( DT, cellVol, J );
}
}
// SAMPLE THE FLOWFIELD TO OBTAIN MACROSCOPIC QUANTITIES
for (N = 0; N < NC; N++) {
cells[N].sample( J );
}
}
} while ( RUN < 1000 ); //LOOP OVER ENSEMBLES
The responsibility of the detailed calculations is contained in the individual
object. The details of the move subroutine, for example, is described in the particle
class, and the details of the collide subroutine is written in the cell class.
53
4.2 Parallel Implementation
Since the code runs independent simulations and then performs ensemble averages
to reduce scatter, it lends itself very nicely to a parallel computing algorithm.
Instead of having one processor do all the independent runs, it is easy to have
dierent computers simultaneously run dierent ensembles, and then average them.
This type of program is called an embarrassingly parallel program and it should
ideally give an eciency of 100%, where eciency is given by:
eciency = 100
speedup
number of processors
, (4.1)
Speedup is the ratio of the CPU time on a serial computer and the CPU time on
a parallel computer.
To achieve this simple parallel processing algorithm, a master/slave routine
is constructed where one processor is considered the master and all the rest are
the slaves [118]. The masters main job is to schedule the processors for work,
organize collected data, and do the nal computing of needed ow properties. The
rest of the processors are the slaves that perform the ensemble runs and send
sampled data back to the master for nal processing. A owchart of the parallel
algorithm is given in Fig. (4.1). The programming here is made possible by using
the Message Passing Interface (MPI) to send messages back and forth between
processors [85, 84].
54
Figure 4.1. A owchart of the parallel algorithm
55
There are numerous benets to using this type of parallel processing. Most
obviously, it is easy to save considerable amounts of time by implementing this
type of program. This program is self-scheduling, so the master computer will
schedule work when needed. Also, on heterogeneous clusters the self-scheduling
will automatically send more work to the fastest processor and the less work to
the slowest processor. Also, the programming for this type of algorithm is quite
simple because there are very few MPI calls.
The code was run run on Mufasa [119], which is a 160-processor cluster at the
Pennsylvania State University and Columbia [120], NASAs fastest supercomputer
with over 10000 processors. Mufasa is a Beowulf cluster running Linux with a
peak speed of 0.4 teraops (oating point operations per second). It has a Dolphin
network for fast communication and 2.8 GHz Athlon processors with about 1 GB of
local memory per processor. Currently, Columbia has a peak speed of 60 teraops
and ranks as the 20th fastest computer in the world [121]. Most of its processors
are Itanium2 1.5 GHz processors with 3.8 GB of local memory per processor.
Despite the ease of this parallel algorithm, using the embarrassingly parallel
technique limits the size of the system that can be simulated due to memory con-
straints of each processor. The memory limitations imply that domain size and
the number of time steps are limited and are machine specic. By limiting domain
length, the largest acoustic wavelength that is possible to simulate with the current
version of code is approximately 2500 cells on Columbia. With the cell size restric-
tion of x 0.5
m
, DSMC is currently limited to situations where Kn 0.0008.
Alternatively, domain decomposition algorithms can be implemented, however the
parallel algorithm is more complicated in this case, requires more interprocessor
communication, and can suer from load balancing issues [122]. This can restrict
the eciency of the parallel algorithm.
In order to illustrate the computational speed of the machines used in this study
as well as the eciency of the algorithm, a one-dimensional version of DSMC was
used to simulate an acoustic wave. Fig. (4.2) shows the wall clock time as a
function of numbers of processors for the same program run on both Columbia
and Mufasa. Computation time on Mufasa in all cases was almost an order of
magnitude larger than on Columbia.
Ideally, a parallel program will scale linearly with the number of processors,
56
Figure 4.2. CPU time on Columbia and Mufasa with respect to number of processors
compared to ideal CPU time
so the speedup of the algorithm on n processors should be n. The speedup on
both Mufasa and Columbia are plotted in Fig. (4.3). On Columbia, performance
is above ideal for more than 32 processors resulting in superlinear speedup. Run-
ning a sequential code on Columbia is a highly inecient use of the computational
resources of the machine. In addition, Columbias architecture allows for shared
memory algorithms as well as distributed memory. In the case of the serial job, even
though a single process was specied, the data was distributed among many pro-
cessors. This increases communication even within serial processing. For greater
than four processors on both Columbia and Mufasa , the speedup is close to being
ideal.
The two processor case on both Columbia and Mufasa is of interest. In this
case, there is one master processor and one slave. Since the slave is the only
processor performing the runs, the added communication with the master makes
it more costly than the serial code on Mufasa. On Columbia, the opposite is true.
This is likely due to the ineciency of Columbia for serial processing. This results
in a superlinear speedup on Columbia for 2 processors as well.
57
Figure 4.3. Speedup on Columbia and Mufasa with respect to the number of processors
compared to ideal speedup
Fig. (4.4) shows the measured eciency versus number of processors. On Mu-
fasa, an eciency of 70% is obtained with only four processors, and it increases
to 98% with 64 processors. However, the eciency drops to 90% for 128 proces-
sors. With only one processor as the master processor and 127 slaves that send
data and messages, the decreased eciency represents a bottleneck in the network
and bandwidth limitations. On Columbia, because of the superlinear speedup,
the eciency is greater than 100% for 32 processors and above as well for two
processors.
Parallel performance is also dependent on system size and network bandwidth.
For large systems, the message sizes are considerable resulting in an increase in
communication time and diminished eciency. Limitations of the network band-
width are seen on Mufasa for larger number of processors. The eciency on Mufasa
for two dierent system sizes is shown in Fig. (4.5). Overall, parallel performance
for larger systems is reduced due to large communication as seen in Fig. (4.5).
Parallel eciency drops to 70% for large systems compared to 90% for smaller
systems.
58
Figure 4.4. Parallel eciency on Columbia and Mufasa with respect to the number of
processors
Despite the eciency of the parallel algorithm, the CPU time for these DSMC
programs can still be fairly large, especially for low Kn situations or for problems
with small deviations from equilibrium.
4.3 Modications to the DSMC algorithm
Because of the large computation time, and slow numerical convergence, several
modications to the DSMC algorithm have been made to simplify the algorithm
and reduce computation costs. These modications are discussed below.
4.3.1 Small deviation from equilibrium
There have been few attempts [123, 124, 125, 126] to modify the DSMC method for
problems in the limit of small deviation from equilibrium, e.g., low Mach number
ow or low acoustic amplitude.
Pan et al. [124] introduced a molecular block modication to DSMC. In
this case, the mass of a big molecule was taken to be the sum of the real mass
59
Figure 4.5. Parallel eciency on Mufasa with respect to the number of processors for
a large and small system
of molecules contained within the molecular block. These molecular blocks were
then used to replace the particles in the DSMC algorithm. Signicant scatter
reduction was reported in the stream velocity and pressure when compared to
traditional DSMC. However, Wang et al. [127] showed later that the molecular
block DSMC method fails as the Mach number increased with increasing mass
of the big molecule. Signicant dierences were found between results from the
molecular block model and the original DSMC method.
A method called DSMC-IP (Information Preservation) was presented by Fan
and Shen [123], and has successfully simulated low-speed gas ow in microchan-
nels. This technique assigns each simulated molecule in the DSMC method two
velocities. One is the molecular velocity used to compute the molecular motion
following the same steps as the DSMC method. The other is called information
velocity. It corresponds to the collective velocity of an enormous number of real
molecules that the simulated molecule represents. Using the information velocity
to compute macroscopic velocity and shear stress may remove the statistical uc-
tuation source inherent in the DSMC method that results from the randomness of
60
the thermal velocity. Despite the amount of scatter reduction, there is increased
complexity and additional memory and computation costs.
Baker and Hadjiconstantinou [128] recently created a variance reduction ap-
proach that focused on ecient methods for evaluating the collision integral. Par-
ticle simulations using this variance reduction technique for solving the Boltzmann
was then presented by Homolle and Hadjiconstantinou [126]. The variance reduc-
tion, achieved by simulating only the deviation from equilibrium, results in a sig-
nicant computational eciency advantage for small deviations from equilibrium
compared to DSMC. Collision rules for the particles that represent the deviation
from equilibrium are be rigorously derived from the new variance-reduced collision
integral. More specically, the proposed method can eciently simulate arbitrar-
ily small deviations from equilibrium at a computational cost that does not scale
with the magnitude of the deviation from equilibrium, while maintaining the basic
algorithmic structure of DSMC.
4.3.2 Low Kn ows modications
There have only been few attempts [129, 122, 74, 130, 65] to use DSMC for problems
where the continuum assumption holds (Kn < 0.05). This is because DSMC
inherently assumes that the cell sizes are on the order of magnitude of the mean
free path of the gas. Typical magnitudes of the mean free path of air in standard
atmospheric conditions are on the order 10
8
m. Therefore, a simulation over large
distances requires enormous time and memory.
Pullin [129] suggested a modication to the DSMC algorithm to extend its ap-
plication to the continuum regime for inviscid, perfect-gas ows. The assumption
of inviscid, perfect-gas ow is equivalent to assuming local thermal equilibrium (Kn
= 0) at every point and at all times [129]. This is analogous to solving the Euler
equations. In his algorithm, Pullin reinitializes the velocity of every molecule to
the local Maxwellian after every iteration, thus ensuring local equilibrium. Inter-
estingly, Pullin [129] ruled out the possibility of using his approach because of the
time penalty associated with the calculation of equilibrium molecular velocities at
each time step.
Sharma and Long [122] have proposed an inexpensive procedure to achieve
61
local thermodynamic equilibrium. In this procedure, one allows enough collisions
per cell per time step to relax the gas to the local Maxwellian. The maximum
number of collisions is a function of the number of particles in the cell but it is
xed at about three-fourths of that number. The procedure of relaxing the gas
to a Maxwellian using collisions is at least about six times more economical than
Pullins [129] suggestion of reinitializing particles to a local Maxwellian velocity at
every time step. A signicant reduction in the computation time by this procedure
makes DSMC a practical solver for low Knudsen numbers.
Hybrid methods combining DSMC and a Navier-Stokes solver are a method
for extending the range of validity of continuum methods to include multi-scaled
problems [131, 132]. The success of hybrid methods depends on how successful
the interface between the continuum method and the particle model is and the
statistical scatter within the DSMC algorithm.
These modications to the DSMC method were pointed out to address some of
the limitations of the algorithm. However, with the computational resources avail-
able, modications to the DSMC algorithm were not necessary for the purposes of
this work.
Chapter 5
Absorption and Dispersion in a
Monatomic Gas
5.1 The speed of sound
The rst recorded measurement of the speed of sound, denoted c, in air was made
in 1635 by Marin Mersenne [133]. The time lapse between the visual sighting of
a cannon ring and the perception of the sound pulse was recorded, and using
the known distance between the cannon and the observer the speed of sound was
calculated to be around 1473 ft/s [133].
Sir Isaac Newton developed a mathematical theory for sound propagation in-
cluding the acknowledgment that sound was created by pressure pulses transmitted
through neighboring uid particles [9, 134]. He dened the speed of sound as:
c
2
=
_
p

_
0
, (5.1)
where the subscript 0 refers to evaluation of the derivative at
0
. However, in his
calculation of the speed of sound, Newton assumes that pressure and density are
directly related by a constant. This relationship, often referred to as Boyles or
Hookes law, results in a calculation of the the speed of sound being 16% lower
than observed measurements of the speed of sound [9].
About 100 years later, Laplace clears up this discrepancy by recognizing the
fact that sound propagation is an adiabatic process with negligible heat ow, and
63
can be written in terms of fundamental thermodynamic properties. In this case,
the relationship between pressure and density for an adiabatic process can be given
by:
p = K

. (5.2)
Therefore, for an adiabatic process the speed of sound is determined to be:
c
2
=
p
0

0
. (5.3)
In the case of a perfect gas, substitution of Eq. (2.16) into Eq. (5.3) results in the
speed of sound given by:
c
2
= 1T/M . (5.4)
The speed of sound is a function of frequency and amplitude. This dependence of
the speed of sound on frequency is called dispersion. The speed of sound in the
limit of low frequency and low amplitude is denoted by a subscript 0, that is, c
0
.
5.2 The absorption of sound
Other mechanisms that govern the propagation of sound in a gas include the ab-
sorption of sound. The nineteenth century was the golden age of classical physics.
At that time, the transfer of acoustic energy into heat by a redistribution of trans-
lational energy among molecules was the only known mechanism for the absorption
of sound and is therefore named classical absorption [135]. For low Kn, classical
loss mechanisms can be described by the Navier-Stokes equations.
The rst published measurement of atmospheric absorption was performed by
Du in 1900 in which an observer stood in a canoe and measured perceived in-
tensity of a whistle source that sounded from a pier above a stream [136]. The
observations concluded that the measured absorption was much larger than the
classical absorption predicted by the Navier-Stokes equations.
More sophisticated experimentation helped unveil the mystery in the 1930s
when Knudsen [137] and Kneser [56] laid the experimental and theoretical basis
for the absorption of sound by molecular relaxation processes. Relaxation losses
associated with the redistribution of translational and internal energy of molecules
64
are the prominent absorption mechanisms for sound in the atmosphere for frequen-
cies less than 10 kHz [135] and will be discussed further in Chapter (6).
However, the redistribution of translational energy introduces dispersion even
in a gas with no internal energy for suciently high Kn. This relaxation process
cannot be modeled using traditional continuum theory based on the Navier-Stokes
equations. A review of the current theory for the absorption and dispersion of
sound in a monatomic, perfect gas will now be given.
5.3 Theory on the absorption and
dispersion of sound in a monatomic gas
The classical linear theory of sound propagation assumes that all acoustic elds
can be written as the sum of an equilibrium value
0
and a small perturbation from
equilibrium

so a harmonic plane wave in the x direction which can be written


in general form as:
=
0
+ Aexp[i(t + kx)] , (5.5)
where A is the acoustic amplitude, which in this case is considered to be small,
is the frequency in rad/s, t is time, and k = + i is the complex propagation
constant where = /c, c is the phase speed, and is the absorption of sound.
The governing continuum equations that describe the perturbation from equi-
librium

are derived from the continuum equations from Sec. (2.4) and will give
the absorption,
cl
, and dispersion,
cl
, due to classical thermal-viscous losses.
The propagation constant for classical losses becomes k
cl
=
cl
+ i
cl
. To write
the conservation equations in terms of the perturbation from equilibrium, start
by considering a system in one dimension. The one dimensional form of the mass
conservation equation, Eq. (2.48), reduces to:

t
+
(U
x
)
x
= 0 . (5.6)
The momentum conservation equation with no external force, Eq. (2.50), reduces
65
in one dimension to:
(U
x
)
t
+
(U
2
x
)
x
+
p
x


xx
x
= 0 , (5.7)
and the conservation of energy equation, Eq. (2.53) in one dimension becomes:

e
t
+
(eU
x
)
x
+
(pU
x
)
x

(U
x

xx
)
x
+
q
x
x
= 0 . (5.8)
The linearized one-dimensional conservation equations are derived from Eqs.
(5.6), (5.7), and (5.8) by assuming small perturbations from equilibrium. Let
p = p
0
+ p

, =
0
+

, u = u
0
+ u

and e = e
0
+ e

then ignoring higher order


terms, the linearized one dimensional conservation equations written in terms of
the perturbation quantities p

, u

and e

are given by:

t
+
0
U

x
x
= 0 , (5.9)

0
U

x
t
+
p

x


xx
x
= 0 , (5.10)

0
e

t
+ p
0
U

x
x

q
x
x
= 0 . (5.11)
The equation of state being used here is the linearized perfect gas law given by
Eq. (2.16). Substituting p = p
0
+ p

, =
0
+

and T = T
0
+ T

into Eq. (2.16)


gives:
p

+ p
0
= (

T
0
+
0
T

+
0
T
0
)
1
M
. (5.12)
Ignoring higher order terms and introducing the low frequency, low amplitude
speed of sound c
0
given by Eq. (5.4), the linearized equation of state for a perfect
gas becomes:

=

c
2
0
p


0
T
0
T

. (5.13)
In the case of a perfect gas, e

= c
v
T

[3]. Therefore Eq. (5.11) can be rewritten


as:
c
v
T

t

1

0
q
x
x
+
p
0

0
U

x
x
= 0 . (5.14)
For a harmonic plane wave, substituting Eq. (5.5) into the linearized conserva-
66
tion equations will give /x = ik
cl
and /t = i. Therefore, the conservation
of mass can be rewritten for a harmonic plane wave as:
i

+ ik
cl

0
U

x
= 0 . (5.15)
The conservation of momentum becomes:
i
0
U

x
+ ik
cl
p

ik
cl

xx
= 0 . (5.16)
For a monatomic gas, the viscous stress tensor is given by Eq. (2.77) which means
that in one dimension,
xx
=
4
3

x
x
. This implies that Eq. (5.16) can be rewritten
as:
i
0
U

x
+ ik
cl
p

+
4
3
k
2
cl
U

x
= 0 . (5.17)
The conservation of energy is given by:
i
0
c
v
T

ik
cl
q
x
+ ik
cl
p
0
U

x
= 0 . (5.18)
The heat ux vector for a hard sphere gas, given by Eq. (2.76), can be introduced
and Eq. (5.18) can be rewritten to produce:
i
0
c
v
T

k
2
cl
T

+ ik
cl
p
0
U

x
= 0 . (5.19)
Eqs. (5.13), (5.15), (5.17), and (5.19) can be combined to create a secular
equation with respect to the classical propagation constant k
cl
which will determine
the dispersion relation for a simple gas [138, 139]. It is shown in Appendix (C) that
the dispersion relation given by the Navier-Stokes equations for a simple gas with
no internal energy written as a function of the propagation constant k
cl
becomes:


2
c
2
0
+
_
i
c
2
0

0
_
4
3
+

c
v
_
+ 1
_
k
2
cl
+

0
c
v
_
4
3c
2
0

_
k
4
cl
= 0 . (5.20)
In the limit of low frequencies, Eq. (5.20) gives the the familiar expression for
the classical absorption coecient
cl
as:

cl
=

2
2
0
c
3
0
_
4
3
+
( 1)
c
p
_
, (5.21)
67
and the phase speed, c
cl
, becomes the adiabatic sound speed given by:
c
cl
=
_
1T
0
M
, (5.22)
The scaled absorption,
cl
/k
0
, and dispersion,
cl
/k
0
, given by Eq. (5.20) are
plotted in Fig. (5.1) along with the low frequency classical absorption coecient
given by Eq. (5.21). k
0
= /c
0
is the acoustic wave number and is inversely
proportional to the acoustic wavelength.
Figure 5.1. Scaled absorption
cl
/k
0
and dispersion
cl
/k
0
predictions given by the
linearized Navier-Stokes equations from Eq. (5.20) are plotted with the low frequency
classical absorption coecient given by Eq. (5.21) [3]
The classical thermal viscous theory given by Eq. (5.20) can accurately describe
the absorption and dispersion in gases with no internal energy for small Kn. This
theory predicts frequency squared dependence for low Kn (low frequency) as seen
in Eq. (5.21), has a maximum value exhibiting relaxation-type behavior, and is
valid within the continuum assumption. When the continuum assumption breaks
down, this theory based on the Navier-Stokes equations can no longer describe
the transport phenomenon correctly, and comparisons between experimental and
computational results with predictions given by Eq. (5.20) for the absorption and
68
dispersion of sound in monatomic gases show signicant deviations at high Kn
due to slow or incomplete translational relaxation [4, 74, 65]. In addition, there is
only an 8.5% dierence between the absorption coecient results of Eqs. (5.20)
and (5.21) at a Kn of 0.05. The dierences between Eqs. (5.20) and (5.4) for the
dispersion of sound is smaller, only 5.2% at a Kn of 0.05. This implies that using
the results of Eq. (5.20) as opposed to Eq. (5.21) may not be advantageous as was
previously thought [4].
In order to account for the deviation from continuum theory at high Kn, there
have been numerous attempts to study sound propagation for high Kn based on
kinetic theory in simple monatomic gases. Most of these attempts use approxima-
tions that replace the Boltzmann collision integral. For example, Wang Chang and
Uhlenbeck [44] studied sound propagation using the method of eigenfunctions. Nu-
merical procedures applied to the eigenfunction method were performed by Pekeris
et al [45]. In Sirovich and Thurber [46] and Loyalka and Cheng [47], sound prop-
agation was studied using a kinetic model description based on the Bhatnagar -
Gross - Krook (BGK) equation described in Sec. (2.5.3) [105]. Marques replaced
the collision integral with a relaxation-time term [48]. Maidanik et al. [49] and
Buckner and Ferziger [50] both use the linearized Boltzmann equation discussed in
Sec. (2.5.2). On the other hand, Zuckerwar uses a phenomenological extension of
the Navier-Stokes equations to describe high Kn behavior [51]. All attempts re-
port that the speed and absorption of sound depend heavily on Kn which deviates
signicantly from traditional continuum theory at high Kn.
Experimental results by Greenspan [4, 52], Meyer and Sessler [53], Maidanik
[54], and Schotter [5] show close agreement for the methods above for Kn < 1, but
for larger Kn, the results vary in degrees of success.
Experimentation at high Kn is dicult [4] and results are sensitive to the
distance between transmitting and receiving transducers [5, 53, 108], especially at
high Kn and will be discussed further in Chapter (7). Collisions of particles with
the receiving transducer introduces a distance dependent propagation constant.
Results taken closer than one mean free path from the transmitting receiver may
also be inuenced by free molecular ow [46]. In order to avoid this dependence, the
receiving transducer of some experimental results was placed at the nondimensional
distance x/u

= 10 [5] where u

is the mean thermal speed of the gas. The


69
theory by Buckner and Ferziger [50, 140] is the only method mentioned above
that accurately accounts for this dependence on distance by correctly solving the
boundary condition problem [141]. This theory uses eigenfunctions and eigenvalues
to solve the linearized Boltzmann equation. The solution by Buckner and Ferziger
has been found to perform well over the whole range for which experimental data
exist. This theory follows classical theory from Eq. (5.20) within the continuum
limit, has a relaxation peak and then decreases to an asymptotic value of 0.2 for
the scaled absorption within the free molecular ow regime [50, 140].
In addition, it has been shown that the acoustic eld is altered by the actual
introduction of receiving transducers, and is dependent on the size of the transducer
face [108]. In this case there is a discrepancy between theoretical results in the free
molecular range.
Large discrepancies between experiment and theory still exist depending on
Kn, distance and the size of transducers used in experimental studies. The DSMC
algorithm is the method best suited to investigate these discrepancies without
relying on linearized theory (continuum or kinetic). This particle method is capable
of describing the acoustic phenomenon for monatomic gases at high Kn where
continuum models fail and is not plagued by the boundary condition problems of
kinetic theory predictions. The validity of the DSMC method for all Kn makes it
the best solution for modeling sound at high Kn.
5.4 DSMC Results for the absorption
and dispersion of sound in a monatomic gas
5.4.1 Simulation approach
DSMC simulations were performed to simulate the sound propagation in gaseous
argon for a wide range of Kn. Acoustic waves were simulated in a one dimensional
simulation domain by creating a piston-like boundary condition at one end of the
domain as described in Chapter (3). In all cases the macroscopic velocity amplitude
of the piston source is 20 m/s. Specular wall reections were implemented at the
opposite side of the simulation domain. In order to minimize computation costs,
the domain length was limited to a few wavelengths. The domain length was 500
70
cells, with each cell initialized with 150 particles per cell. Each simulated particle
represents one real particle, so the ratio of real to simulated particles, F
num
, is
one in each case. Gaseous argon was simulated with molecular weight M = 39.94
kg/kmole and cross-sectional diameter = 3.66 10
10
m.
The cell size, x was required to be less than a half of a mean free path so
the error in the transport coecients is less than 2.5% [114, 115]. The variation in
Kn was obtained by specifying the frequency of oscillation of the piston boundary
condition and by maintaining the number of cells per wavelength constant at 100.
This required the cell size to vary from 1/2 of a mean free path to 1/200 of a
mean free path. The cell size implies that the domain ranges from 6.7 10
5
m
to 1.6 10
5
m. Because of the cell size, the number of particles, the number of
points per wavelength, and the ratio of real to simulated particles, the mean free
path also varies as Kn increases.
The time step was taken to be at least an order of magnitude smaller than the
mean collision time and is on the order of picoseconds for each case. Care was
taken to ensure that the time step remained smaller than the acoustic period of
oscillation.
Calculations were performed at a temperature of 273 K. Each case was initial-
ized in thermal equilibrium so that T
x
= T
y
= T
z
.
5.4.2 Results
The scaled absorption
cl
/k
0
as a function of Kn is shown for 273 K in Fig. (5.2).
DSMC results were computed by tracking the maximum pressure amplitude as a
function of distance and computing an exponential line of best t for extracting
the absorption coecient to a distance governed by the nondimensional distance
x/u

= 10 [5]. DSMC results are also plotted against theoretical predictions


given by the imaginary part of Eq. (5.20) for the absorption coecient and also
the classical low frequency absorption coecient given by Eq. (5.21). DSMC
results are also plotted with experimental results by Greenspan [4] and Schotter
[5]. Both experimental and DSMC results agree quite well with continuum theory
for low Kn. However, large deviations from continuum theory are seen for high
Kn, as expected due to the breakdown of the continuum. The simulation results
71
approach the free molecular ow limit of 0.2 at high Kn [5, 50]. Mid Kn results
near the translational relaxation frequency depend more heavily on simulation or
experimental setup.
The scaled dispersion
cl
/k
0
as a function of Kn is shown for 273 K in Fig.
(5.3). DSMC results were computed by tracking the zero crossings of the acoustic
pressure as a function of distance and computing the slope of the line of best t for
the phase speed to a distance governed by the nondimensional distance x/u

= 10
[5]. DSMC results are also plotted against theoretical predictions given by the real
part of Eq. (5.20). Experimental results for the dispersion in argon is only available
by Greenspan [4] and is plotted with DSMC results. As with the absorption, both
experimental and DSMC results agree quite well with continuum theory for low
Kn. However, large deviations from continuum theory are seen for high Kn, as
expected due to translational relaxation. Comparisons with experimental results
are quite good over the entire Kn range.
Figure 5.2. Scaled absorption
cl
/k
0
in argon for 273 K. DSMC results (red circles)
compared to experimental results by Greenspan [4] (green triangles) and Schotter[5]
(blue squares) and continuum theory predictions results from Eq. (5.20) [3]
72
Figure 5.3. Scaled dispersion k/k
0
in argon for 273 K. DSMC results (red circles)
compared to experimental results by Greenspan [4] (green triangles) and continuum
theory predictions results from Eq. (5.20) [3]
5.4.2.1 Nonequilibrium
Signicant deviations from continuum theory occur at high Kn due to slow or
incomplete translational relaxation. The physics of molecular relaxation implies
that when a system that starts in equilibrium is subject to a sound source, there
is a time delay between the exchange in energy between translational modes. At
high Kn when translational relaxation eects are more evident, the system is
in a higher state of nonequilibrium. The details of this relaxation process are
shown in Fig. (5.4) where the overall translational temperature and the x, y, and
z components of the translational temperature computed using Eqs. (2.17) and
(2.18) respectively are shown for a frequency of 23 MHz corresponding to Kn = 2
at 273 K. Despite starting in equilibrium, the temperatures associated with the
x, y, and z components of the translational temperature in this case are quite
dierent. Very little energy has relaxed from the x component into the other
components.
In contrast, the system is still in equilibrium for a frequency of 230 MHz cor-
73
responding to Kn = 0.02 as seen in Fig. (5.5).
Figure 5.4. Translational nonequilibrium eects for Kn = 2.0 at 273 K in argon.
The ability to investigate the measure of nonequilibrium in a system is unique
to DSMC which makes it the method best suited to investigate nonequilibrium
eects, in particular at high Kn. This particle method is capable of describing
the details of translational relaxation in a monatomic gases at high Kn where
continuum models fail. The validity of the DSMC method for all Kn allows for
the study of nonequilibrium eects, even within the continuum assumption which
signies the strengths of the DSMC algorithm. Nonequilibrium eects as a function
of amplitude will be discussed further in Chapter (7).
74
Figure 5.5. Translational nonequilibrium eects for Kn = 0.02 at 273 K in argon.
Chapter 6
Absorption and Dispersion in a Gas
With Internal Energy
6.1 Introduction
The physical properties that govern the absorption and dispersion of sound in a
gas include classical losses associated with the transfer of acoustic energy into heat
as was discussed in Chapter (5) and relaxation losses associated with the redistri-
bution of internal energy of molecules. Current theories regarding relaxation losses
due to rotational and vibrational energy will now be given.
6.2 Absorption and dispersion from a simple
relaxation process
For any relaxation process, there is a nite time required to establish equilibrium
between the translational and internal states. In the case of a single relaxing
internal degree of freedom, the translational and internal temperatures, T
tr
and
T
int
, respectively, satisfy the relaxation equation:
T
int
t
=
1

r
(T
tr
T
int
) , (6.1)
where
r
is the relaxation time.
76
This relaxation process introduces dispersion,
r
= /c
r
, which takes the form:
_

r
k
0
_
2
= 1 s
f
2
fr
1 +
_
f
fr
_
2
. (6.2)
and absorption,
r
that can be written as:

r
k
0
=
k
0
s
2
r
f
fr
1 +
f
2
f
2
r
. (6.3)
s is the relaxation strength and f
r
is the relaxation frequency or the frequency of
maximum absorption due to the relaxation process [55, 3, 142]. The relaxation
strength is given by the equation:
s =
1C

p
(C

v
+ C

)
, (6.4)
where C

is the heat capacity of the relaxing mode, and the superscript denotes
conditions that are well above the relaxation frequency.
The relaxation frequency f
r
can be written in terms of the relaxation time
as:
f
r
= 1/(2
r
) . (6.5)
The relaxation time can be written as a function of the relaxation collision number,
Z, dened in Chapter (3) as:

r
=
C

p
Z
C
0
p

, (6.6)
where = u

/
m
is the collision frequency, u

is the average thermal speed in an


equilibrium gas, and
m
is the mean free path as dened in Chapter (2). The
superscript 0 denotes conditions that are well below the relaxation frequency.
The absorption due to translational relaxation was presented in Sec. (5.3).
The relaxation behavior derived from the Navier-Stokes equations is concealed by
the complexity of the dispersion relaxation of Eq. (5.20). However, Zuckerwar [51]
uses a phenomenological approximation to write Eq. (5.20) in terms of a relaxation
time. Therefore, Eq. (5.20) can be cast in the form of Eqs. (6.2) and (6.3) using
Zuckerwars approximations.
77
6.3 Current theories including rotational
relaxation
The absorption and dispersion by rotational relaxation,
rot
and
rot
= /c
rot
respectively, are usually written in the form of Eqs. (6.2) and (6.3) and can usually
be modeled using a single relaxation time [135]. In the case of rotational relaxation
in a diatomic gas with 2 degrees of rotational freedom at low temperatures where
rotation is the only active internal mode, the heat capacity of the relaxation mode,
C

is 1 which is based on Eq. (2.22). For frequencies higher than the relaxation
frequency, the rotational degrees of freedom do not contribute to the specic heats
and therefore

= 5/3. Therefore, C

p
= 51/2 and C

v
= 31/2. Therefore, the
relaxation strength s in the case of rotational relaxation in a diatomic gas is 4/25
Greenspan [6] presented classical theoretical expressions for the combined ab-
sorption and dispersion due to classical (translational) and rotational relaxation.
It was found that if the gas is a Becker gas, with a Prandlt number c
p
/ of 3/4,
then Eq. (5.20) is factorable and has an explicit solution [6]. The Prandlt number
for common gases is shown in Table (6.1). The Becker gas assumption is an ade-
quate approximation for most gases with the exception of monatomic gases. Using
the Becker gas assumption, Greenspan has shown that it is possible to write the
combined absorption,
cr
, and dispersion,
cr
, due to translational and rotational
relaxation as:

cr
=

cl

rot
k
0
+

rot

cl
k
0
, (6.7)
and

cr
=

cl

rot
k
0


cl

rot
k
0
. (6.8)

cl
and
cl
result from Eq. (5.20).
At low frequencies, Eq. (6.7) reduces to
cr
=
cl
+
rot
and all absorption
mechanisms are additive. In the low frequency limit,
rot
can be rewritten as [135]:

rot
=
2
2
f
2
p
0
c
( 1)1
1.25C
0
p
Z
rot
. (6.9)
The scaled absorption including rotational relaxation,
cr
/k
0
, and dispersion
including rotational relaxation,
cr
/k
0
, according to Eqs. (6.7) and (6.8) is plotted
78
Table 6.1. The Prandlt number for common gases
gas Prandtl number
Ne 0.66
Ar 0.67
N
2
0.71
CH
4
0.74
O
2
0.72
CO
2
0.75
H
2
0.71
in Figs. (6.1) and (6.2) respectively. The results are given for relaxation collision
numbers of Z
rot
=1, 10, 100, and 1000 in a nitrogen-like gas with molecular weight
28.01 at 0

C and plotted with predictions by the Navier-Stokes equations, Eq.


(5.20), and low frequency assumptions given by Eqs. (5.21) and (5.22).
The relaxation frequency increases with decreasing collision number, as seen in
Figs. (6.1) and (6.2). The relaxation peak is visible for collision numbers of 100
and 1000, but for Z
rot
< 100, classical thermal-viscous losses dominate for all Kn.
In addition, the relaxation frequency for collision numbers less than 5 are beyond
the continuum assumption and thus Eqs. (6.7) and (6.8) are not expected to hold
under such circumstances since they assume Navier-Stokes solutions for classical
losses.
Based upon experimental results, the rotational relaxation collision number
Z
rot
is a function of temperature. For most gases used in this study, the rotational
relaxation collision number can be determined by the empirical equation [13, 143]:
Z
rot
= 61.1 exp(16.8/T
1/3
) . (6.10)
Eqs. (6.5), (6.6), and (6.10) indicate that the relaxation frequency increases with
decreasing temperature.
79
Figure 6.1. Scaled absorption for collision numbers Z
rot
=1, 10, 100, and 1000 in a
nitrogen-like gas at 0

C from Eqs. (6.7) and (6.3) [3, 6]


Figure 6.2. Scaled dispersion for collision numbers Z
rot
=1, 10, 100, and 1000 in a
nitrogen-like gas at 0

C from Eqs. (6.8) and (6.2) [3, 6]


80
6.3.1 Other theories for rotational relaxation
Greenspans rule given by Eqs. (6.7) and (6.8), which is by far the most popular in
describing translational and rotational relaxation absorption, is only satised for
a Becker gas and has no rigorous kinetic theory justication [62]. In addition, it is
only valid for a single relaxing degree of freedom. However, McCormack generalizes
Greenspans rule for more than one relaxational degree of freedom [63].
The rst attempt at predicting sound behavior in polyatomic gases based on
a kinetic theory description was made by Wang Chang and Uhlenbeck [57]. They
extended the Chapman-Enskog treatment described in Sec. 2.5.2 for the case of
polyatomic gases. Monchick et al. [58] and Mason and Monchick [59] investigated
the case for more than one internal degree of freedom based on a kinetic theory
approach. Hanson and Morse [60] and Hanson et al. [61] use an extension of the
kinetic theory for a monatomic gas presented by Sirovich and Thurber [46] for a
polyatomic gas by using a modied BGK equation. The kinetic theory method
used by McCormack and Creech [62] is a moment expansion of the distribution
function from the Boltzmann equation which are of special physical signicance.
Banankhah and Loyalka [64] use the equations of Wang Chang and Uhlenbeck to
numerically solve the boundary value problem of diuse boundaries.
Comparisons between the theoretical predictions of Eqs. (6.7) and (6.8) and
experimental results show good agreement within the continuum assumption [52].
However, for Kn > 0.05 the results show poor agreement. Sutherland and Bass
[144] use an empirical adjustment to account for high Kn deviation from the-
ory. The kinetic theory solutions presented by references [57] - [64] show better
agreement for higher Kn than Greenspans rule. The solution by Banankhah and
Loyalka correctly solves the boundary value problem and shows the best agreement
to experimental results by Greenspan [64].
Despite the fact that Greenspans method for combining classical thermo-
viscous and rotational relaxation fails for high Kn, it is still the most popular
way of describing the absorption and dispersion for rotational relaxation [144].
81
6.4 Current theories including vibrational
relaxation
It is widely known that vibrational relaxation losses play an important role in
sound propagation at audible frequencies in the Earths atmosphere [145]. The
introduction of water vapor to a dry air mixture has been shown to shift the
vibrational relaxation frequencies of oxygen and nitrogen higher in frequency with
increasing humidity. This happens through complicated vibrational energy transfer
pathways during intermolecular collisions. The absorption and dispersion due to
rotational relaxation as presented in Sec. 6.3 was simplied since in most cases
only one relaxation time is necessary in describing the relaxation process. In that
case, the absorption could be modeled simply using Eq. (6.3) and the dispersion
could be simply modeled using Eq. (6.2).
However, in the case of vibrational relaxation, the exchange of vibrational en-
ergy between translational modes occurs at very dierent rates. This is because
the discrete energy levels for vibration are spaced further apart than for rotation.
In addition, the exchange of vibrational energy between other vibrational modes of
colliding molecules complicates the theory in such a way that there is no good way
to go from measured relaxation times of specic relaxation reactions to a general
formulation for the relaxation frequency. In the absence of vibration-vibration en-
ergy transfer, it is possible to write the relaxation frequency with a single relaxation
time for each reaction as was done with rotational relaxation [135].
There are many experimental results in binary mixtures for specic relaxation
times of dierent relaxation reactions [135, 146, 147, 148]. But there is no way
a priori to determine which reactions will be most important in the study of the
absorption of sound in a gas mixture. For the case of vibrational relaxation as a
function of humidity in the Earths atmosphere, empirical equations were deter-
mined for the relaxation frequency as a function of humidity.
In addition, the relaxation frequency for vibration is a complicated function of
temperature. Experimental work in shock tubes has given the relaxation time for
many diatomic gases as a function of temperature [7, 8]. The results show that
the vibrational relaxation frequency increases with increasing temperature which
is the opposite of the behavior seen in rotational relaxation [149]. The relaxation
82
frequency for vibration in nitrogen and oxygen based on experimental work [7, 8]
is plotted in Fig. (6.3) as a function of temperature. The relaxation frequency for
vibration is many orders of magnitude lower than for rotational relaxation.
Figure 6.3. Vibrational relaxation frequency for oxygen and nitrogen as a function of
temperature [7, 8]
Once the relaxation frequency is determined for the vibrational relaxation re-
action i, then the absorption and dispersion due to the i
th
relaxation process
vib,i
can be determined using Eqs. (6.2) and (6.3).
The total absorption coecient including rotational and vibrational relax-
ation losses is then written [144, 9] as:
=
cr
+

vib,i
, (6.11)
where
cr
is usually given by Eq. (6.7).
DSMC simulations investigating the validity of the theoretical models presented
above for gases with internal energy will now be given.
83
6.5 Implementation in DSMC for a gas with
rotational energy for multiple collision
numbers
6.5.1 Simulation approach
To show the ability of DSMC to model relaxation phenomena, simulations were
performed including rotational internal energy for a range of relaxation collision
numbers. DSMC simulations were performed to simulate the sound propagation
in a nitrogen-like gas for a wide range of Kn. Acoustic waves were simulated in a
one dimensional simulation domain by creating a piston-like boundary condition
at one end of the domain as described in Chapter (3). In all cases the macroscopic
velocity amplitude of the piston source is 20 m/s. Specular wall reections were
implemented at the opposite side of the simulation domain. In order to minimize
computation costs, the domain length was limited to a few wavelengths. Each
cell is initialized with 50 particles per cell. Each simulated particle represents one
real particle, so the ratio of real to simulated particles, F
num
, is one in each case.
Gaseous nitrogen was simulated with molecular weight M = 28.01 kg/kmole and
cross-sectional diameter = 3.78 10
10
m.
The cell size, x was required to be less than a half of a mean free path.
Therefore, cell sizes were restricted to x 0.5
m
. The variation in Kn was
obtained by specifying the frequency of oscillation of the piston boundary condition
and by maintaining the number of cells per wavelength constant at 100. This
required the cell size to vary from 1/2 of a mean free path to 1/200 of a mean free
path. The cell size implies that the domain ranges from 4 10
6
m to 4.8 10
5
m. Because of the cell size, the number of particles, the number of points per
wavelength, and the ratio of real to simulated particles, the mean free path also
varies as Kn increases.
The time step was taken to be at least an order of magnitude smaller than the
mean collision time and is on the order of picoseconds for each case. Care was
taken to ensure that the time step remained smaller than the acoustic period of
oscillation.
Calculations were performed at a temperature of 273 K. Each case was initial-
84
ized in thermal equilibrium so that T = T
tr
= T
rot
. Because nitrogen is a diatomic
gas with one axis of rotational symmetry, the gas was modeled using two rotational
degrees of freedom for all cases.
6.5.2 Results
The scaled absorption coecient, /k
0
is plotted as a function of Kn for relax-
ation collision numbers 1, 5, 40 and 200 in Figs. (6.4), (6.5), (6.6) and (6.7)
respectively. Results are shown with theoretical predictions given by Eq. (6.3) [3]
and the combination of rotational relaxation and classical losses [6] given the colli-
sion numbers used. The frequency range investigated spans the Knudsen number
range of 0.0008 Kn 5. It is expected that DSMC results would follow theory
within the continuum assumption and approach the free molecular limit for the
scaled absorption of 0.2 for high Kn. The rotational relaxation peak is evident
in the case with Z = 200, but in all the other cases, classical thermal-viscous
losses become more important so the relaxation peaks are less dened. There is
evidence of a relaxation mechanism for the case of Z = 40, but for Z = 5 and
Z = 1, rotational relaxation occurs beyond the continuum assumption and si-
multaneously with translational relaxation. Despite being beyond the continuum
assumption, the scaled absorption does depend on the relaxation collision number
for Kn > 0.05.
For ease of comparison, the DSMC results for all collision numbers are plotted
together in Fig. (6.8) with the theoretical predictions given by Eq. (6.3) [3] and
classical thermal-viscous losses from the Navier-Stokes dispersion relation of Eq.
(5.20). In the cases of Z = 40 and Z = 200 the rotational mode of the molecules
is in its frozen state and does not contribute to the specic heats for frequencies
associated with Kn > 0.2. The DSMC results for these frequencies are nearly
identical for Z = 40 and Z = 200. For Z = 1 and Z = 5, the scaled absorption is
more dependent on Kn for Kn > 0.05 indicating that there is a unique combination
of rotational relaxation and translational relaxation at these frequencies. However,
all cases for the relaxation collision number approach the free molecular limit for
Kn > 2.0.
85
Figure 6.4. Scaled absorption for relaxation collision number of 1. DSMC simulations
(points) are plotted with continuum theory for the rotational relaxation given by Eq.
(6.3) (dashed line) [3] and total absorption given by Eq. (6.7) (solid line) [6]
Figure 6.5. Scaled absorption for relaxation collision number of 5. DSMC simulations
(points) are plotted with continuum theory for the rotational relaxation given by Eq.
(6.3) (dashed line) [3] and total absorption given by Eq. (6.7) (solid line) [6]
86
Figure 6.6. Scaled absorption for relaxation collision number of 40. DSMC simulations
(points) are plotted with continuum theory for the rotational relaxation given by Eq.
(6.3) (dashed line) [3] and total absorption given by Eq. (6.7) (solid line) [6]
Figure 6.7. Scaled absorption for relaxation collision number of 200. DSMC simulations
(points) are plotted with continuum theory for the rotational relaxation given by Eq.
(6.3) (dashed line) [3] and total absorption given by Eq. (6.7) (solid line) [6]
87
Figure 6.8. DSMC results for the scaled absorption with relaxation collision numbers
of 1, 5, 40, and 200. DSMC simulations (points) are plotted with continuum theory for
rotational relaxation (dashed line) [3]
6.6 Implementation in DSMC for a gas with
rotational and vibrational energy
Due to the additional importance of vibrational relaxation eects on the absorption
and dispersion of sound, the relative importance of the various internal energy
models with DSMC as a function of Kn for gases with rotational and vibrational
energy has also been investigated. Classical and quantum vibrational models that
were introduced in Chapter (3) have been implemented to investigate their eect
on acoustics as a function of temperatures.
6.6.1 Simulation approach
Acoustic waves were again simulated in a one dimensional simulation domain by
creating a piston-like boundary condition at one end of the domain. In all cases
the macroscopic velocity amplitude of the piston source is 20 m/s. Specular wall
88
reections were implemented at the opposite side of the simulation domain. The
domain length was 500 cells, with each cell initialized with 50 particles per cell
on average. Each simulated particle represents one real particle, so the ratio of
real to simulated particles, F
num
, is one in each case. Results for varying Kn
were simulated in a nitrogen-like gas with xed cross-section (molecular weight
M = 28.01 kg/kmole and cross-sectional diameter = 3.78 10
10
m).
The variation in Kn was obtained by specifying the frequency of oscillation
of the piston boundary condition and by maintaining the number of cells per
wavelength constant at 100, and varying the cell size from 1/2 of a mean free path
to 1/200 of a mean free path. Because of the cell size, the number of particles, the
number of points per wavelength, and the ratio of real to simulated particles, the
mean free path also varies as Kn increases. The cell size implies that the domain
ranges from 4 10
6
m to 4 10
5
m.
The time step was taken to be at least an order of magnitude smaller than the
mean collision time and is on the order of picoseconds for each case. Care was
taken to ensure that the time step remained smaller than the acoustic period of
oscillation.
In each case presented, the rotational collision number, Z
rot
, was chosen to be
5 and the vibrational collision number, Z
vib
was chosen to be 200. While these
choices for rotational and vibrational collision numbers do not reect tempera-
ture dependence as seen by Eq. (6.10) or even the correct order of magnitude
for vibration as seen in experimental results for nitrogen [7, 8], the goal was to
investigate the internal energy models and their validity for the DSMC method.
The rotational collision number was chosen to be 5 based upon experimental mea-
surements of nitrogen at room temperature [143]. Under the worst conditions, the
total number of collisions performed over the entire simulation was calculated to
be only 60000 given the small scale of the system. In order to simulate vibrational
relaxation eects during the simulated time, the vibrational collision number was
chosen to be 200 given the goal of modeling the dierence between internal energy
models. Therefore, the choice of collision numbers was based on the size of the
system, so that the relaxation eects could be monitored more easily given the
frequency, temperature, and Kn ranges investigated.
Calculations were performed at temperatures of 273 K, 2000 K, and 4000 K in
89
order to monitor the temperature dependence of the excitation of the vibrational
mode of nitrogen. For each case, the rotational degree of freedom for nitrogen was
modeled classically, while the vibrational degree of freedom was either modeled
using the classical model, the quantum model, or modeled without vibration. The
classical model for rotation is a good assumption in nitrogen because of the closely
spaced rotational energy levels. However, in some gases, most notably diatomic
hydrogen, the discrete energy levels for rotation are spaced further apart than most
other diatomic gases [149]. In this case, rotational relaxation can not be modeled
using a single relaxation time and more complex quantum theory is needed to
describe the rotational relaxation behavior. Experimental results for temperatures
greater than 2000 K indicate that there may be more than one rotational relaxation
frequency in hydrogen [149].
The characteristic temperature of nitrogen was specied as 3371 K, which for
low freestream temperatures allows for a single vibrational energy level to ade-
quately model the vibrational temperature. It has been shown by Bird that a
single vibrational level is adequate to model higher temperature ows [1]. An ex-
ample of this vibrational modeling at higher temperatures is shown in Fig. (3.5).
For all approaches, nitrogen was modeled using two rotational degrees of freedom
with each degree of freedom represented by a single square term.
Each case was initialized in thermal equilibrium so that T
tr
= T
vib
= T
rot
, but
since molecular relaxation is a nonequilibrium process, deviation from equilibrium
is expected. Section 6.6.2.1 describes this in further detail.
Dissociation of diatomic nitrogen in the high-temperature/low-pressure systems
was found to be negligible for the conditions at which the simulations were run. An
independent simulation was completed using DSMC containing nitrogen molecules
at a temperature of 4000 K. This simulation represented the system shown in
Fig. (6.14) in Section 6.6.2 which was predicted to have the greatest likelihood
for dissociation. A reaction model was employed and the nitrogen molecules were
given the opportunity to dissociate. Both the number of collision and dissocia-
tion events were recorded over the entire simulation time. It was determined that
approximately 11,000 collisions were encountered, on average, between each disso-
ciation. Based on the dissociation rate, the total simulation time and the pressure,
the simulations in Fig. (6.14) are predicted to have less than one molecule undergo
90
dissociation over the entire simulation.
6.6.2 Results
6.6.2.1 Nonequilibrium
The physics of molecular relaxation imply that when a system that starts in equi-
librium is subject to a sound source, there is a time delay between the exchange
in energy between translational and internal modes. The details of this relaxation
process for classical vibration are shown in Fig. (6.9) where the temperatures
associated with the translational, rotational, and classical vibrational modes are
computed for a frequency of 402 MHz corresponding to Kn = 0.02 at 273 K. The
relaxation time for the vibrational model is considerably longer than that for ro-
tation and is evident by the minimal disturbance in the vibrational temperature.
Similar nonequilibrium eects are seen at higher temperatures.
At high Kn when slow translational and rotational relaxation eects are more
evident, the system is in a higher state of nonequilibrium. Relaxation is a very slow
process at high Kn where the frequency of oscillation is well above the relaxation
frequencies for rotation and vibration. The degree of nonequilibrium can be noted
in Fig. (6.10) for a frequency of 57 MHz corresponding to Kn = 1.0 at 273 K.
Despite starting in equilibrium, the temperatures associated with the translational,
rotational, and classical vibrational modes in this case are considerably dierent.
Very little energy has relaxed from the translational mode into the internal modes
where the internal modes are in their frozen state.
91
Figure 6.9. Nonequilibrium eects for Kn = 0.02 at 273 K with classical vibration
model
Figure 6.10. Nonequilibrium eects for Kn = 1.0 at 273 K with classical vibration
model.
92
At low temperatures the quantum vibrational model will exhibit almost no
vibrational activity as a large majority of diatomic nitrogen molecules will remain
in the ground state due to the molecules high characteristic temperature. In
contrast, the classical model allows for a continuous distribution of vibrational
energy determined by the capacity of the molecule and energy available. Fig.
(6.11) shows the fraction of molecules in the excited state in each cell for the
quantum vibration model at the temperatures 273 K, 2000 K, and 4000 K for
Kn = 0.02. As the temperature increases, more molecules are in the excited state,
as noted in the gure.
Figure 6.11. Fraction of molecules in the excited state for Kn = 0.02 at 273 K (red
square), 2000 K (green triangle) and 4000 K (blue circle).
6.6.2.2 Absorption as a function of temperature
The scaled absorption /k
0
as a function of Kn is shown for 273 K, 2000 K, and
4000 K in Figs. (6.12), (6.13), and (6.14), respectively. DSMC results were com-
puted by tracking the maximum pressure amplitude as a function of distance and
computing an exponential line of best t for extracting the absorption coecient to
a distance governed by the nondimensional distance x/u

= 10 [5]. DSMC results


are also plotted against theoretical predictions given by the imaginary part of Eq.
(5.20) for classical thermal-viscous losses, the classical low frequency absorption
93
coecient given by Eq. (5.21), and the combined losses including rotational relax-
ation given by Eq. (6.7). The results for the quantum, classical, and vibrationless
models are shown for a range of Kn in each case. Large deviations from continuum
theory are seen for high Kn, as expected due to the breakdown of the continuum.
The simulation results approach the free molecular ow limit of 0.2 in each case
[5]. In each temperature case, dierences between the models are small for low
Kn.
Figure 6.12. (Scaled absorption in nitrogen for 273 K. DSMC results (symbols) com-
pared to continuum theory predictions given by Eqs. (5.20), (5.21), and (6.7) (lines)
[3, 6].
Due to the fundamental dierences between the models, simulations using con-
tinuous and discrete vibrational models at low temperatures are expected to pro-
duce dierent results. The classical harmonic oscillator model assumes a fully
active vibrational mode while the quantum harmonic oscillator does not. The
quantum harmonic oscillator model is a more realistic model at low temperatures.
For evenly spaced energy levels the quantum and classical models become identical
in the limit of high temperatures.
In the 273 K case the quantum and vibrationless congurations yield similar
results for all Kn. This is because at low temperatures the fraction of molecules
94
Figure 6.13. Scaled absorption in nitrogen for 2000 K. DSMC results (symbols) com-
pared to continuum theory predictions given by Eqs. (5.20), (5.21), and (6.7) (lines)
[3, 6].
in excited states is nearly negligible and the vibrational energy is very close to
zero, as in the vibrationless case. The classical vibration results are lower than the
quantum and vibrationless cases for high Kn and higher than the quantum and
vibrationless cases for low Kn. In this case, the dierences arise because of dierent
nonequilibrium eects in the classical vibration case which are a function of the
choice in vibration relaxation collision number. In addition, the theory plotted in
the gures given by Eqs. (5.20), (5.21), and (6.7) do not take into account the
eects of vibrational relaxation and are only valid for one relaxation process.
As the temperature increases, the excited state for quantum vibration becomes
more populated and behaves more like the classic harmonic oscillator. Therefore,
the dierences between continuous and discrete vibrational models become less,
as is shown in Figs. (6.13) and (6.14). By 4000 K, all three internal energy
congurations yield similar results within the error where the quantum vibration
is highly populated.
Due to the stochastic nature of DSMC, there is an intrinsic degree of scat-
ter in the simulation results. Scatter is reduced by averaging many independent
95
Figure 6.14. Scaled absorption in nitrogen for 4000 K. DSMC results (symbols) com-
pared to continuum theory predictions given by Eqs. (5.20), (5.21), and (6.7) (lines)
[3, 6].
ensembles and random error is estimated from variance in repeated runs. As the
temperature increases, the amount of scatter increases due to the increase in molec-
ular speed. The amount of random error was found to have no signicant eect
on the results shown for 273 K. However, for 2000 K and 4000 K the error is
more signicant. Error bars indicating one standard deviation in the calculated
scaled absorption values are given in the gures after calculations for the scaled
absorption were made on ten independent data sets.
6.6.2.3 Dispersion as a function of temperature
The scaled dispersion /
0
as a function of Kn is shown for 273 K, 2000 K, and 4000
K in Figs. (6.15), (6.16), and (6.17), respectively. DSMC results were computed
by tracking zero crossings of the acoustic pressure as a function of distance and
computing a line of best t for extracting the phase speed to a distance governed
by the nondimensional distance x/u

= 10 [5]. DSMC results are also plotted


against theoretical predictions given by the real part of Eqs. (5.20) for classical
96
dispersion, (6.2) for rotational dispersion and (6.8) for total dispersion. The results
for the quantum, classical, and vibrationless models are shown for a range of Kn
in each case. Large deviations from continuum theory are seen for high Kn, as
expected.
In contrast to the absorption cases, the dispersion seems to be less dependent
of internal energy congurations for all temperatures. Similarly to absorption, the
classical vibration conguration consistently produces results that are lower than
the vibrationless and quantum vibration cases. In addition, the intrinsic amount of
scatter in the results does not increase signicantly as the temperature increases.
Figure 6.15. Scaled dispersion in nitrogen for 273 K. DSMC results (symbols) compared
to continuum theory predictions given by Eqs. (5.20), (6.2), and (6.8) (lines) [3, 6].
97
Figure 6.16. Scaled dispersion in nitrogen for 2000 K. DSMC results (symbols) com-
pared to continuum theory predictions given by Eqs. (5.20), (6.2), and (6.8) (lines)
[3, 6].
Figure 6.17. Scaled dispersion in nitrogen for 4000 K. DSMC results (symbols) com-
pared to continuum theory predictions given by Eqs. (5.20), (6.2), and (6.8) (lines)
[3, 6].
Chapter 7
The Eect of Amplitude
Continuum or kinetic theory approaches typically assume that the amplitude of
the acoustic source is small. When the acoustic amplitude is large, nonlinear
eects are only well understood within the continuum assumption. Nonlinear
phenomena such as shock formation, harmonic generation, and shock coalescence
can be studied using continuum equations within the continuum approximation,
but for high Kn the eects of amplitude are unknown when translational relaxation
and free molecular ow become important. Because of the uncertainty of the
relative importance of relaxation eects and nonlinearity on the propagation of
sound at high Kn, the dependence of amplitude in a monatomic gas has been
considered here.
7.1 Simulation approach
Acoustic waves were generated in the simulation domain by creating a piston
boundary condition at one end of the domain. The macroscopic velocity amplitude
of the piston-like source ranged from 5 m/s to 40 m/s. Specular wall reections
were implemented at the opposite side of the simulation domain. The amplitude
was increased to 150 m/s in Section (7.6) for a strong nonlinearity study. Results
for varying Kn were simulated in gaseous argon as hard sphere molecules at 273
K (molecular weight M = 39.94 kg/kmol and hard sphere diameter 3.66 10
10
m). In order to minimize computation costs, the domain length was limited to a
few wavelengths. The domain length was 500 cells, with each cell initialized with
99
150 particles per cell. Each simulated particle represents one real particle, so the
ratio of real to simulated particles, F
num
, is one in each case.
The variation in Kn was obtained by specifying the frequency of oscillation
of the piston boundary condition and by maintaining the number of cells per
wavelength constant at 100 for most cases. The cells per wavelength was increased
to 400 in Section (7.5) to ensure sucient resolution for Fourier transform analysis.
This required the cell size to vary from 1/2 of a mean free path to 1/200 of a mean
free path. The cell size implies that the domain ranges from 6.7 10
5
m to
1.6 10
5
m. Because of the cell size, the number of particles, the number of
points per wavelength, and the ratio of real to simulated particles, the mean free
path also varies as Kn increases.
The time step was taken to be at least an order of magnitude smaller than the
mean collision time and is on the order of picoseconds for each case. Care was
taken that the time step remained smaller than the acoustic period of oscillation.
Each case was initialized to start in thermal equilibrium.
7.2 Breakdown of the propagation constant
Continuum theory for the absorption and dispersion of sound in a monatomic
gas presented in Chapter (5) assumes small deviation from equilibrium, that is,
small acoustic amplitude. Linear theory predicts that not only does the acoustic
amplitude decay exponentially as a function of distance, but that the propaga-
tion constant, k, is independent of the distance from the source. However, as
mentioned previously, collisions of particles with the receiving transducer create
a dependence on distance for the propagation constant which has been conrmed
analytically [50, 140], experimentally [5] and computationally [74]. In order to
avoid this dependence, the receiving transducer of some experimental results was
placed at the nondimensional distance x/u

= 10 [5] where u

is the mean thermal


speed of the gas.
To illustrate the dependence of the propagation constant on distance, the same
data from DSMC simulations is considered for a frequency of 23 MHz corresponding
to Kn = 2 in argon at 273 K with the data t restricted for dierent values of x.
The maximum pressure amplitude with the distance restricted to x < 10u

/ is
100
shown in Fig. (7.1) as a function of distance. The macroscopic velocity amplitude
of the piston source is 40 m/s in this case. The results indicate that the wavelength
and attenuation coecient attain constant values for x < 10u

/. The exponential
t extrapolated in the region x > 10u

/ deteriorates very quickly as shown in Fig.


(7.2).
Figure 7.1. Maximum pressure amplitude (Pa) in argon for Kn = 2, 40 m/s amplitude.
DSMC results (points) compared to restricted exponential best t curve for x < 10u

/
(solid line)
The dependence on distance is most prominent in high Kn, high amplitude
conditions, and the magnitude of the deterioration depends on the amplitude of
the piston source. For small amplitudes, the propagation constant dependence on
distance is not severe even for highly nonequilibrium situations. A frequency of 23
MHz corresponding to Kn = 2 in argon at 273 K has also been considered with a
velocity amplitude of 5 m/s and is shown in Fig. (7.3). In this case, the maximum
pressure amplitude tted for all values of x ts well with an exponential model
indicating that high amplitude eects play an important role in Figs. (7.1) and
(7.2).
101
Figure 7.2. Maximum pressure amplitude (Pa) in argon for Kn = 2, 40 m/s amplitude.
DSMC results (points) compared to exponential best t curve for x > 10u

/ (solid line)
Figure 7.3. Maximum pressure amplitude (Pa) in argon for Kn = 2, 5 m/s amplitude.
DSMC results (points) compared to exponential best t curve for x > 10u

/ (solid line)
102
In addition, a case with small Kn and high amplitude performed well for all
values of x as shown in Fig. (7.4). In this case, a frequency of 230 MHz corre-
sponding to Knudsen number of 0.02 is considered in argon at 273 K with velocity
amplitude of 40 m/s. The results comparing high amplitude studies at Kn = 2
from Fig. (7.2), and Kn = 0.02 from Fig. (7.4) indicate that the propagation
constant behavior as a function of distance is a unique combination of nonlinear
eects and nonequilibrium eects only at high Kn.
Figure 7.4. Maximum pressure amplitude (Pa) in argon for Kn = 0.02, 40 m/s am-
plitude. DSMC results (points) compared to exponential best t curve for x > 10u

/
(solid line)
In all cases, restricting the dataset to distances of x < 10u

/ is consistent with
experimental studies [5] and produces the best results for nonexponential decay
for all values of Kn and amplitudes.
103
7.3 Nonequilibrium eects as a function
of amplitude
Despite the fact that all cases have been initialized to be in thermal equilibrium
so that T
tr
= T
trx
= T
try
= T
trz
, translational nonequilibrium eects are evident
at high Kn and the magnitude of nonequilibrium is a function of amplitude. Fig.
(7.5) shows the translational nonequilibrium, T
trx
/T
tr
, for the Kn = 2, 40 m/s
case at a snapshot in time. There is almost a 15% dierence between T
trx
and T
tr
indicating strong nonequilibrium. In contrast, the amount of nonequilibrium in
the case with Kn = 2 and an amplitude of 5 m/s is an order of magnitude smaller
than the case with Kn = 2 and 40 m/s amplitude as seen in Fig. (7.6).
Figure 7.5. T
trx
/T
tr
for Kn = 2, 40 m/s amplitude.
104
Figure 7.6. T
trx
/T
tr
for Kn = 2, 5 m/s amplitude.
Figure 7.7. T
trx
/T
tr
for Kn = 0.02, 40 m/s amplitude.
105
Figure 7.8. T
trx
/T
tr
for Kn = 0.02, 5 m/s amplitude.
Even at large amplitudes, nonequilibrium eects for lower Kn (Kn = 0.02, 40
m/s amplitude in this case) are an order of magnitude smaller than for high Kn,
high amplitudes, as shown in Fig. (7.7). On the other hand, nonequilibrium eects
for low Kn, low amplitude are within the noise oor of the system as shown in
Fig. (7.8) signifying that the system is in equilibrium. Therefore, it can be inferred
that high amplitude eects, regardless of Kn, cause the system to be in a higher
state of nonequilibrium as seen in Figs. (7.5) and (7.7) than lower amplitudes.
In addition, translational relaxation increases the magnitude of nonequilibrium in
higher Kn systems shown in Figs. (7.5) and (7.6).
7.4 Absorption and dispersion as a function of
amplitude
The scaled absorption /k
0
as a function of Kn is shown in Fig. (7.9) for macro-
scopic velocity amplitudes of 5, 10, 20, and 40 m/s. DSMC results are shown with
the Navier-Stokes theoretical predictions and the low frequency approximation for
106
the classical absorption coecient given by Eqs. (5.20) and (5.21) respectively.
The results for the dierent piston amplitudes are shown for a range of Kn in
each case. The scaled absorption data was computed within the range x < 10u

/
for each case to avoid nonexponential behavior described in Section (7.2). Large
deviations from continuum theory are seen for high Kn, as expected. The dif-
ferences between the scaled absorption for dierent amplitudes are small for high
and low Kn. For Kn near the translational relaxation frequency, the amount of
absorption depends heavily on the amplitude. Despite the small dierences in the
scaled absorption coecient at low and high Kn, nonlinear eects are evident in
the larger amplitudes as seen in Fig. (7.10) with signicant wave steepening for
the Kn = 0.02, 40 m/s amplitude case, for example.
Figure 7.9. Amplitude dependence on the scaled absorption in argon at 0

as a function
of Kn. DSMC results (points) compared to continuum theory given by Eq. (5.20) (line)
The scaled dispersion k/k
0
as a function of Kn is shown in Fig. (7.11) for
macroscopic velocity amplitudes of 5, 10, 20, and 40 m/s. DSMC results are
shown with the Navier-Stokes theoretical predictions given by the real part of Eq.
(5.20). The results for the dierent piston amplitudes are shown for a range of
Kn in each case. The scaled dispersion was calculated by tracking the maximum
107
Figure 7.10. Wave steepening at Kn = 0.02 and 40 m/s amplitude
of the acoustic pressure as a function of distance and computing the slope of the
line of best t for the phase speed within the range x < 10u

/. Large deviations
from continuum theory are seen for high Kn, as expected due to translational
relaxation. At low and high Kn, DSMC results are quite dependent on amplitude.
The low Kn behavior is a result of classical nonlinear eects. The phase speed of
a progressive nonlinear wave is given by [10]:
c = c
0
+
NL
U
0
, (7.1)
where
NL
is the coecient of nonlinearity and is dened by
NL
= (+1)/2, where
is the ratio of specic heats, U
0
is acoustic velocity amplitude, and c
0
is the low
frequency, low amplitude speed of sound. Thus, the phase speed is amplitude
dependent and is evident when considering the maximum pressure amplitude as a
function of distance at low Kn as seen in Fig. (7.11).
The zero crossings of the acoustic pressure should not be amplitude dependent
within the continuum assumption and should travel at c
0
. The results for the scaled
dispersion that were computed using the zero crossings of the acoustic pressure as
108
Figure 7.11. Amplitude dependence on the scaled dispersion in argon at 0

as a function
of Kn based on the maximum pressure amplitude. DSMC results (points) compared to
continuum theory given by Eq. (5.20) (line)
opposed to the maximum pressure amplitude as a function of distance are shown
in Fig. (7.12). In this case, the amplitude dependence at low Kn is very small in
contrast to the results plotted in Fig. (7.11).
In general, the results shown in Figs. (7.11) and (7.12) behave quite dierently.
The results from Fig. (7.11) show amplitude dependence at low and high Kn
whereas the results from Fig. (7.12) indicate the opposite behavior. The dispersion
results calculated from the zero crossings of the acoustic pressure behave in a
similar manner as a function of amplitude to the absorption shown in Fig. (7.9)
with amplitude dependence near the translational relaxation frequency.
The amplitude dependence near the relaxation frequency can be explained by
investigating the number of collisions per molecule for each system. The number of
collisions performed in each cell is shown in Fig. (7.13) for a 40 m/s amplitude case
and 5 m/s amplitude case at a frequency of 73 MHz corresponding to Kn = 0.2
after a time of 1 nanosecond. As shown in Fig. (7.13) the number of collisions per
molecule is signicantly higher in the high amplitude case than the low amplitude
case. More collisions in the higher amplitudes will bring the system to a higher
109
Figure 7.12. Amplitude dependence on the scaled dispersion in argon at 0

as a function
of Kn based on the zero crossings of the acoustic pressure. DSMC results (points)
compared to continuum theory given by Eq. (5.20) (line)
state of equilibrium. This in turn will cause lower amounts of absorption. This
dependence on amplitude is the reverse of what is seen in the continuum when
there is extra attenuation due to nonlinearity [150].
7.5 Harmonic generation
The distance x = 10u

/ was calculated to be smaller than the traditional shock


formation distance, x, for amplitudes 5, 10, and 20 m/s. For the 40 m/s amplitude,
the distance x = 10u

/ was no more than one and a half times the shock formation
distance. The shock formation distance is derived from the second order one-
dimensional wave equation and is given by the equation [10, 9]:
x =
1
(
NL
k
0
)
, (7.2)
where = U
0
/c
0
is the acoustic Mach number. However, the shock formation
distance is derived from the continuum approximation and is not necessarily valid
110
Figure 7.13. Number of collisions performed in each cell for a 40 m/s amplitude (dashed
line) and 5 m/s amplitude (solid line) at Kn = 0.2 after a time of 1 nanosecond
for high Kn.
To nondimensionalize the spatial scale, the scaled shock formation distance is
commonly written:

NL
=
x
x
, (7.3)
which implies that the wave shocks initially at the value = 1.
An implication of shock formation is that higher harmonics develop with in-
creasing propagation distance from the source. The amplitude of the fundamental
reduces while the amplitudes of the higher harmonics get larger. The Fourier
component amplitudes of the rst three harmonics were compared to the analyti-
cal Fubini solution [10, 9] and Burgers equation as a function of nondimensional
shock propagation distance given by Eq. (7.3) and nondimensional distance given
by x/u

= 10 for a frequency of 58 MHz corresponding to Kn = 0.005, 40 m/s am-


plitude in Fig. (7.14). The Fubini solution gives the Fourier component amplitude
B
n
for the nth harmonic as:
B
n
=
2
n
J
n
(n
NL
) , (7.4)
111
where J
n
is the Bessel function of the rst kind of order n.
The Fubini solution is valid for
NL
< 1 and does not include losses due to
thermo-viscous absorption and dispersion. For the case of plane wave propagation
in a thermo-viscous medium, an analytical solution of Burgers equation includes
both nonlinear and classical losses. Mendousses solution of Burgers equation is
given by [9]:
p
p
0
=
4

n=1
(1)
n+1
I
n
_

2
_
ne
n
2
/
sin(nt

)
I
0
_

2
_
+ 2

n=1
(1)
n
I
n
_

2
_
e
n
2
/
cos(nt

)
, (7.5)
where is the Goldberg number, t

= t x/c
0
is the retarded time, and I
n
are
modied bessel functions. The Goldberg number is dened as:
=
1
x
0
, (7.6)
where x is the shock formation distance and
0
is the low frequency classical
absorption coecient[10]. For Goldberg number values of > 1 nonlinear eects
dominate and shock formation is imminent. For < 1 absorption dominates over
nonlinearity and dissipative eects attenuate the acoustic signal enough to keep
shock waves from forming. However, it should be noted that since the Goldberg
number is dened by continuum parameters and the low frequency approximation
of the classical absorption coecient, the Goldberg number is not well dened
for high Kn situations. In addition, it also may not be well suited for frequencies
where the absorption due to a relaxation process dominates since by denition,
the Goldberg number is a function of the classical absorption coecient.
The solution to Burgers equation, given by Eq. (7.5), outputs a time waveform,
so the computed solution will need to be broken down into their Fourier component
amplitudes for comparison.
Because the Fubini theory does not include thermo-viscous losses, the Fubini
solution overpredicts the component amplitudes for all harmonics in comparison
to DSMC results. On the other hand, DSMC results compare very well with the
Burgers equation solution for each harmonic.
High amplitude eects beyond the continuum assumption are poorly under-
112
Figure 7.14. Fourier component amplitude for Kn = 0.005, 40 m/s amplitude com-
pared to Fubini solution given by Eq. (7.4) and Burgers equation given by Eq. (7.5)
[9, 10].
stood and should not be expected to follow the Fubini solution. The Fourier
component amplitude for Kn = 2, 40 m/s is shown in Fig. (7.15). In comparing
Figs. (7.14) and (7.15) it is seen that for high Kn the higher harmonics initially
are more dominate in the Kn = 2 case than in the Kn = 0.005 case, despite faster
decay of the fundamental.
In contrast, the Fourier component amplitude for the rst 3 harmonics for high
Kn and low amplitude is shown in Fig. (7.16). In this case there is no harmonic
generation. The rate of decay of the fundamental frequency is approximately the
same as in the high amplitude case which explains the equivalent absorption values
that were shown in Fig. (7.9).
113
Figure 7.15. Fourier component amplitude for Kn = 2, 40 m/s amplitude.
Figure 7.16. Fourier component amplitude for Kn = 2, 5 m/s amplitude.
114
7.6 Shock coalescence
Other high amplitude eects can also be simulated using the DSMC method such
as shock coalescence. During shock coalescence, small shocks are overtaken by large
shocks because of the large dierences in the phase speed which can be determined
from Eq. (7.1). With large amplitudes and small absorption, larger shock waves
travel much faster than smaller shock waves and the waves coalesce before the wave
dissipates. The progression of a highly nonlinear wave for Kn = 0.02 is shown in
Figs. (7.17) - (7.20). The macroscopic velocity amplitude is 150 m/s. The rst
and second shocks are beginning to coalesce in Fig. (7.18) and by Fig. (7.20) the
by Fig. (7.20) there is a large overpressure that spans six periods of oscillation.
Figure 7.17. Shock coalescence for Kn = 0.02 at 6 nanoseconds
115
Figure 7.18. Shock coalescence for Kn = 0.02 at 15 nanoseconds
Figure 7.19. Shock coalescence for Kn = 0.02 at 21 nanoseconds
116
Figure 7.20. Shock coalescence for Kn = 0.02 at 46 nanoseconds
7.6.1 Coalescence at high Kn
A form of coalescence can be seen at high Kn even at lower amplitudes. The
progression of a 40 m/s wave for Kn = 2 is shown in Figs. (7.21) - (7.24). The
overpressures dominate the underpressures as time progresses. Despite high atten-
uation, only overpressures are visible in Fig. (7.24) and the wave has coalesced.
There are very few collisions at Kn = 2 and so the molecular speed is much higher
in the overpressure than in the underpressures. Similarly to the coalescence at
low Kn, large amplitude waves overtake the small amplitude waves and the waves
coalesce resulting in signicant overpressures.
117
Figure 7.21. Coalescence for Kn = 2 at 42 nanoseconds
Figure 7.22. Coalescence for Kn = 2 at 63 nanoseconds
118
Figure 7.23. Coalescence for Kn = 2 at 74 nanoseconds
Figure 7.24. Coalescence for Kn = 2 at 84 nanoseconds
Chapter 8
DSMC Applications: Planetary
Acoustics
The Cassini-Huygens spacecraft is the largest interplanetary, international space-
craft ever built. Its mission is to perform an in-depth study of the diverse phenom-
ena of the Saturn system, including the planet itself, its rings and its largest moon,
Titan. The orbiter, Cassini, boasts cutting-edge instruments capable of collecting
sophisticated data for 250 scientists in 17 countries. It was also designed to piggy-
back the Huygens probe all the way to Titans orbit, and to be able to eject the
probe properly. The orbiter also was designed to forward to Earth the information
it receives from Huygens during the probes descent through the thick atmosphere
of Titan [151]. The Huygens probe descended to the surface of Titan in early 2005,
capturing data on the structure, composition and climate of Titans atmosphere.
Acoustic sensors recorded ambient sounds in hopes of capturing thunder in the
moons thick, dense and cloud-lled atmosphere [152, 153]. Other planetary mis-
sions have also included acoustic sensors in hopes that they would transmit sound
samples back to Earth. In the 1980s, the Russian-built Venera mission lasted 120
minutes on Venus and survived long enough to record a possible thunder event
[154]. But it wasnt again until 1999 when a low-cost microphone was constructed
for the ill-fated Mars Polar Lander mission [155] that an acoustic sensor was cre-
ated for planetary exploration. While the Mars Polar Lander lost contact with
Earth shortly after its descent into the Martian atmosphere and never recovered,
there is hope that a future mission will include more acoustic experiments. With
120
Table 8.1. Atmospheric conditions at the surface on Earth, Mars and Titan
planet Earth Mars Titan
temperature (K) 300 200 95
pressure (atm) 1.0 0.007 1.5
molecular composition N
2
(70%), O
2
(20%) CO
2
(95%) N
2
(95%), CH
4
(5%)
the recent success of the Huygens probe, there has been renewed interest in further
exploring the acoustic environments of the other planets in the solar system.
The planetary environments of Earth, Mars and Titan provide a wide range of
atmospheric conditions to study the propagation of sound, which is very sensitive
to changes of pressure, temperature and molecular composition. On these planets
(while Titan is a moon, it will be referred to as a planet), atmospheric pressures
range from 0.007 atm on Mars to 1.5 atm on Titan and temperatures range from
90 K on Titan to 300 K on Earth. As on Earth, the majority of Titans atmo-
sphere is made up of nitrogen, while Mars harsh atmosphere is made up mostly of
carbon dioxide. The atmospheric composition on the surface of the three planets is
summarized in Table (8.1) [11, 12, 14]. The vastly dierent atmospheres of Earth,
Mars and Titan present a wide range of options for exploring extra terrestrial
sound propagation.
Despite the fact that several space missions were implemented with acoustic
sensors, little work has been done to quantify sound propagation in planetary
atmospheres based on non-empirical models. Even less has been done to study
sound waves of nite amplitude where nonlinear eects could be present. Notable
contributions include a study of absorption and dispersion near the Martian surface
based on semi-empirical methods [13], an overview of the acoustic environment on
Mars [156], and a comparison of atmospheric linear acoustics on Mars, Titan,
Venus and Earth using an eective wavenumber approach [157].
Particle simulations of sound propagation investigating the eects of amplitude,
relaxation and altitude on Earth, Mars and Titan will now be presented.
121
8.1 Simulation approach
DSMC simulations were performed to simulate the sound propagation on Earth,
Mars and Titan. On each planet, acoustic waves were simulated in a one di-
mensional domain by creating a piston-like boundary condition at one end of the
domain as described in Chapter (3). In all cases the macroscopic velocity ampli-
tude of the piston source is 20 m/s. Specular wall reections were implemented
at the opposite side of the simulation domain. In order to minimize computation
costs, the domain length was limited to a few wavelengths.
Molecules were simulated as hard sphere particles. The cell size, x was re-
quired to be less than a half of a mean free path. The variation in Kn was obtained
by varying the acoustic wavelength from 100 to 2500 cells.
The time step was taken to be at least an order of magnitude smaller than the
mean collision time and is on the order of picoseconds for each case. Care was
taken to ensure that the time step remained smaller than the acoustic period of
oscillation.
Calculations were performed at a temperature of 273 K. Each case was initial-
ized in thermal equilibrium so that T = T
tr
= T
rot
.
Details on the implementations for each planet are given below. Atmospheric
properties for all planets were derived from from public-access general circulation
models [11, 158, 159, 12, 160, 14].
8.2 Results and Discussion
8.2.1 Earth
The Earths atmosphere close to the surface contains roughly 78% nitrogen, 20.95%
oxygen, 0.93% argon, 0.038% carbon dioxide, and trace amounts of other gases
(most notably, water vapor). The average atmospheric pressure at sea level is
about 101000 Pa with average temperatures near 300 K.
However, pressure and temperature are also functions of altitude above the
surface. Fig. (8.1) gives the temperature and pressure proles as a function of
altitude up to 300 km [11].
122
Figure 8.1. Temperature and pressure proles as a function of altitude above Earths
surface [11]
In an eort to model vibrational relaxation eects that are so important to
sound propagation in the Earths atmosphere, DSMC simulations were performed
in Earth-like conditions. The absorption of sound due to a simple relaxation pro-
cess given by a single relaxation time is presented in air at atmospheric conditions
(p = 101000 Pa and T = 273 K).
Because the addition of water vapor in the atmosphere decreases the relaxation
time, two sets of simulations were run to represent dry and humid air on Earth.
The relaxation time is proportional to the vibrational collision number, Z
vib
which
is given to be 200 for the dry air case and Z
vib
= 40 for the humid air case.
The rotational relaxation numbers for nitrogen and oxygen should be very
close at atmospheric temperatures. Based on experimental values, the rotational
relaxation number, Z
rot
, for these species is given by Eq. (6.10) [13].
While the choice of vibrational collision numbers is not necessarily physically
realistic, the choice of collision numbers was based on the size of the system. In this
case, relaxation eects could be investigated within the computational restrictions
that DSMC imposes on the domain size.
123
The scaled absorption coecient, /k
0
is plotted as a function of Kn for the
dry air case in Fig. (8.2) and the humid air case in Fig. (8.3). Results are
shown with theoretical predictions for vibrational relaxation [3] and the combina-
tion of vibrational relaxation, rotational and classical losses [6] given the collision
numbers used. The frequency range investigated spans the Knudsen number range
of 0.0008 Kn 0.05, which is within the continuum assumption. Therefore,
it is expected that DSMC results would follow theory. The relaxation peak is
evident in the dry air case, but in the humid air case, classical thermal-viscous
losses become more important so the relaxation peak is less dened as was also
seen in Chapter (6) for rotational relaxation. The relaxation peak does increase in
Knudsen number when decreasing the collision number (increasing humidity) as
shown in Figs. (8.2) and (8.3).
Figure 8.2. Scaled absorption for dry air on Earth. DSMC simulations (points) are
plotted with continuum theory for the vibrational relaxation given by Eq. (6.3) (green)
and total absorption given by Eq. (6.7) (red)
In addition, dry air earth conditions have also been modeled for frequencies
above the vibrational relaxation frequencies where classical losses dominate. A
frequency of 230 MHz corresponding to Kn = 0.02 was simulated in a gas mixture
of nitrogen, oxygen, and argon at atmospheric conditions (p = 101000 Pa and T =
124
Figure 8.3. Scaled absorption for humid air on Earth. DSMC simulations (points)
are plotted with continuum theory for the vibrational relaxation given by Eq. (6.3)
(green) and total absorption given by Eq. (6.7) (red)
273 K). DSMC results for the acoustic pressure at a point in time are compared to
the predicted amplitude dependence determined from the Navier-Stokes dispersion
relation of Eq: (5.20) in Fig. (8.4). Comparisons between the DSMC results and
theory are quite good as expected.
8.2.2 Mars
The molecular composition of the Martian atmosphere is well established by previ-
ous missions and ground based observations. The most prominent constituent near
the surface is carbon dioxide (95.3%). There are also minor amounts of nitrogen
(2.7%) and argon (1.6%), with trace amounts of oxygen and water vapor [161].
Mars thin atmosphere has a surface pressure that varies greatly throughout
the planet, which supports large winds that create planet-wide dust storms. The
average pressure at the surface ranges from 500-700 Pa which is only about 0.7%
of that on Earth at sea level. Also, the temperature on Mars in cooler than Earth,
ranging from 200 K to 300 K which can result in frozen carbon dioxide on the
125
Figure 8.4. DSMC results (solid line) for the acoustic pressure on Earth for Kn =
0.02 compared to predicted amplitude dependence determined from the Navier-Stokes
derived absorption coecient from Eq. (5.20)(dashed line)
polar caps.
Data is available for atmospheric conditions extending over 200 km above the
Martian surface [159]. Temperature and pressure proles are plotted as a function
of altitude in Fig. (8.5). DSMC simulation conditions were drawn from these
models for a variety of altitudes [158, 159, 12].
Because of the low temperatures on Mars, vibrational energy is not active for
frequencies above the relaxation frequency and does not contribute to the spe-
cic heat. The relaxation frequency for carbon dioxide, the primary atmospheric
constituent, is below 100 Hz under Martian conditions [13]. However, DSMC sim-
ulations were performed for frequencies far above the relaxation frequency, so only
the rotational internal energy of the molecules needs to be considered.
The rotational relaxation numbers for carbon dioxide, nitrogen and oxygen
should be very close at Martian temperatures [13] and can be given given by Eq.
(6.10).
The dierences in atmospheric conditions between Earth and Mars result in
the low amplitude, low frequency speed of sound on Mars being 66% lower than
126
Figure 8.5. Temperature and pressure proles as a function of altitude above the Mars
surface [12]
on Earth [162]. This coupled with the low ambient density results in an acoustic
impedance which is two orders of magnitude smaller than on Earth. This two
order of magnitude dierence implies that not only would it take 100 times the
particle velocity to get the same acoustic pressure as on Earth, but that pressure
wave signals will be 20 dB weaker on Mars [157] for the same acoustic pressure
ignoring attenuation.
Model predictions by Bass and Chambers [13] and Petculescu and Lueptow
[157] give the absorption of sound on the surface of Mars to be at least 100 times
larger than on Earth. This is dominated by classical thermal-viscous losses for
frequencies within the continuum limit and larger than 10 KHz. For frequencies
smaller than 10 KHz, absorption is dominated by vibrational relaxation losses from
the relaxation of the doubly degenerate bending mode of carbon dioxide [13]. This
allows the absorption and dispersion to be written in terms of a single relaxation
time so the theory is not complicated by complex energy transfer pathways as on
Earth. In addition, Bass and Chambers use the low frequency approximation to
the classical absorption coecient.
127
DSMC results for the maximum acoustic pressure amplitude as a function of
distance is plotted in Fig. (8.6) for Kn = 0.02. This corresponds to a frequency
of 1.3 MHz, which is signicantly above the vibrational relaxation frequency of
carbon dioxide and within the continuum assumption. At this frequency, classical
thermal-viscous absorption dominates. Simulations were performed at 700 Pa and
200 K. In this case, a mixture of carbon dioxide, nitrogen, argon and oxygen
(water vapor and other trace gases were omitted in the simulations) was considered.
DSMC results are plotted against theoretical predictions presented by Bass and
Chambers. Comparisons between the DSMC results and theoretical predictions
are quite good.
Figure 8.6. DSMC results for the acoustic pressure amplitude as a function of distance
on Mars for Kn = 0.02 compared to theoretical predictions [13]
Rotational relaxation at Martian temperatures occurs simultaneously with trans-
lational relaxation given the low relaxation collision numbers given by Eq. (6.10).
For frequencies corresponding to Kn > 0.05, the time delay between the exchange
in energy between translational and internal modes becomes more noticeable, cre-
ating a higher state of nonequilibrium. At high Kn the frequency of oscillation
is well above the relaxation frequencies for rotation. DSMC results showing the
degree of nonequilibrium is plotted in Fig. (8.7) where the temperatures associated
128
with the translational and rotational modes are computed for a frequency of 130
MHz corresponding to Kn = 2. Despite starting in equilibrium, the temperatures
associated with the translational and rotational modes in this case are consider-
ably dierent. In this case, slow translational and rotational relaxation eects are
more evident. Very little energy has relaxed from the translational mode into the
rotational mode where the rotation mode of the molecules is in its frozen state.
Figure 8.7. Nonequilibrium eects on Mars showing rotational (dashed line) and trans-
lational (solid line) temperatures for Kn = 2
8.2.3 Titan
Scientists have tried for decades to penetrate the thick, hazy atmosphere of Titan
with a variety of telescopes, but were only able to obtain vague hints at the shape
of the surface below. Recently, however, the Cassini-Huygens mission unveiled a
great deal about Titan and its atmosphere [163, 151]. Titans landscape resembles
a younger, colder Earth which has an atmosphere of 90 to 97 percent nitrogen, with
at least a dozen of other organic compounds, including methane, argon, hydrogen,
ethane and water, implying that Titans atmosphere is chemically active.
At Titans surface, its pressure is 1.6 atm and its temperature of 90 K is much
129
colder than that on Earths surface. The atmosphere extends 500 km above the
surface, and the temperature, pressure and molecular composition are a complex
function of altitude [160, 163]. Temperature and pressure proles are plotted as a
function of altitude in Fig. (8.8) [14]. In comparing Fig. (8.8) to Fig. (8.1), there
is only a 6 order of magnitude change in pressure over 500 km in altitude on Titan
as opposed to over 10 order of magnitude change in pressure over only 300 km on
Earth. In addition, the molecular composition as a function of altitude does not
change very much. Nitrogen and methane are still the prominent constituents at
high altitudes. Hydrogen is the next most prominent gas in the upper atmosphere,
but is present in trace amounts (0.1%) [160].
Figure 8.8. Temperature and pressure proles as a function of altitude above Titans
surface [14]
For the DSMC simulations of Titan, a mixture of 95% nitrogen and 5% methane
(trace gases were omitted in our analysis) was considered.
Similar to Mars, because of the low temperatures on Titan, the vibrational
energies of nitrogen or methane are not active and do not contribute to the specic
heat for the frequencies under consideration. Therefore, we only need to take into
consideration the rotational internal energy of the molecules.
130
The rotational relaxation number for nitrogen is again based on experimental
values given by Eq. (6.10). Experimental values for the rotational relaxation
number for methane are not available at temperatures considered here, thus the
value of 5 is used as a best guess.
Because of the large concentrations of nitrogen on both Titan and Earth, the
molecular weights are comparable between the two planets. However, with the low
temperatures and molecular composition on Titan the speed of sound is 60% of
the speed of sound on Earth and 88% of the speed of sound on Mars. In addition,
due to an order of magnitude dierence in acoustic impedance between Earth and
Titan, pressure wave signals in Titans dense atmosphere would be 10 dB higher
than on Earth for the same acoustic pressure making Titan acoustically responsive
[157]. This implies that it is easier to produce and sustain high-amplitude signals,
possibly leading to signicant nonlinear eects.
Predictions by Petculescu and Leuptow give the absorption of sound to be 10
times smaller on Titan than on Earth because of the dense pressures and low
temperatures [157]. At these low temperatures, the molecules are in their frozen
state and absorption is therefore dominated by classical thermal-viscous losses.
Because of the low absorption and high acoustic impedance, nonlinear waves can
travel a long distance before being absorbed by the atmosphere.
Given the order of magnitude dierence between the absorption coecient on
Titan and Earth and the similarity between the coecient of nonlinearities on
both planets, the Goldberg number, given by Eq. (7.6), on Titan is an order of
magnitude higher than on Earth for the same frequency and amplitude. As an
example, DSMC simulations were performed on Earth and Titan for a frequency
of 230 MHz and acoustic Mach number = 0.14 and are shown in Figs. (8.9) and
(8.10). This corresponds to a Kn = 0.042 on Earth and Kn = 0.01 on Titan, well
within the continuum approximation. For this frequency, classical thermal-viscous
losses will dominate on both planets, so the Goldberg number is a good measure
of the importance of nonlinearity. The Goldberg number is calculated to be 3.8 on
Titan and 0.95 on Earth. This implies, and Fig. (8.10) suggests, that nonlinearity
dominates for this frequency and amplitude on Titan as evidenced by signicant
wave steepening of the waveform. In contrast, Fig. (8.9) shows little, if any, wave
steepening and signicant attenuation on Earth, demonstrating the dominance of
131
absorption over nonlinearity in this case.
Figure 8.9. Kn = 0.042 waveform on Earth with = 0.14. Absorption dominates
nonlinearity with little or no nonlinear eects visible.
8.2.4 Vertical proles
On all three planets, atmospheric pressure decreases approximately exponentially
with altitude as shown in Figs. (8.1), (8.5) and (8.8), and the relationship between
temperature and altitude varies among the dierent atmospheric layers. Close to
the surface, the atmosphere can be assumed to be continuum for low frequencies
and can be treated as a perfect gas. However, the continuum assumption gradually
breaks down as altitude increases and diusion and vertical transport become more
important. In order to fully understand the acoustic environment as a function of
altitude, particle methods are necessary when the Knudsen number reaches 0.05
in order to capture nonequilibrium and high Kn eects.
The Knudsen numbers for several frequencies as function of altitude are plot-
ted for Earth, Mars and Titan in Figs. (8.11), (8.12) and (8.13), respectively.
On Earth, all frequencies over 100 Hz reach the continuum limit by 100 km in
altitude. Similarly on Mars, the continuum limit is reached at approximately 90
132
Figure 8.10. Kn = 0.01 waveform on Titan with = 0.14. Signicant wave steepening
can be observed.
km in altitude for the same frequencies. However, on Titan, because of its dense
atmospheric pressure, the continuum assumption is still valid for frequencies less
than approximately 1000 Hz at all altitudes. This implies that particle method
solutions are required when investigating the absorption and dispersion of sound
for frequencies above 100 Hz and 100 km on Earth and Mars because of the devi-
ation from continuum theory as was seen in Chapters (5) - (7). However, particle
methods are necessary only for frequencies above 1 MHz at 100 km altitude on
Titan.
Previous work for studying sound propagation as a function of altitude is given
for Earth up to 160 km [144] and for Earth, Venus, Mars and Titan up to 120 km
[157]. Sutherland and Bass use a theory by Greenspan [6] to model translational
and rotational relaxation and an empirical adjustment to account for the devia-
tion to theory at high Kn. Diusion losses and vibrational relaxation losses are
also accounted for based on estimates of the relaxation frequencies of the primary
molecular constituents as a function of altitude. Temperature and molecular com-
position dependence was considered in mean atmospheric conditions, including:
humidity, viscosity, and the specic heat ratio. Rotational relaxation and classical
133
Figure 8.11. Knudsen number as a function of altitude on Earth for frequencies of 100
Hz, 1000 Hz, 10000 Hz, and 1000000 Hz
thermal viscous losses were found to be the prominent loss mechanisms at altitudes
above 90 km.
In the predictions given by Petculescu and Lueptow [157], an eective wave
number approach developed by Dain and Lueptow [164] is used to model relaxation
phenomenon. In this technique, the linearized continuum conservation equations
in addition to temperature relaxation equations are used to form an eigenvalue
problem. The relaxation equations include kinetic theory results for collision rates
and quantum mechanics predictions for transition probabilities. The result can be
formed into an eective wave number for determining the acoustic absorption and
dispersion due to a relaxation process. Petculescu and Lueptow use this eective
wavenumber theory in conjunction with the low frequency classical absorption to
describe the acoustic properties in planetary environments. While this method
may prove to be computationally eective, the assumptions of using the linearized
continuum equations in addition to the limit of low frequencies limits its usefulness.
In addition, translational and rotational relaxation, which has been shown to be
important in high Kn situations was not considered. Thus, care must be taken
134
Figure 8.12. Knudsen number as a function of altitude on Mars for frequencies of 100
Hz, 1000 Hz, 10000 Hz, and 1000000 Hz
when applying this eective wavenumber approach as a function of altitude.
DSMC results for the scaled absorption /k
0
for a frequency of 70 MHz and
altitude of 25 km were computed for Earth, Mars and Titan based upon general
circulation model data. Given the atmospheric conditions of each planet, this
choice of frequency results in values of Knudsen numbers that span a large range.
Mars, with its thin and transparent atmosphere results in the largest Knudsen
number of 16, high above the continuum breakdown and into the free molecular ow
regime. On Earth, the Knudsen number is smaller, but still above the continuum
limit at 0.47, and Titan with its dense, thick atmosphere results in a Knudsen
number of 0.01 at this frequency. The scaled absorption /k
0
for each planet
at an altitude of 25 km is shown in Fig. (8.14) and compared to Navier-Stokes
predictions for classical thermal-viscous losses [139]. Deviation from Navier-Stokes
is seen for results on Earth and Mars given the high Knudsen numbers associated
with the frequency of 70 MHz as expected. Results for Titan, however, agree quite
well with predicted values as expected for the Kn associated with this simulation.
The results for Earth and Mars approach the free-molecular limit for the scaled
135
Figure 8.13. Knudsen number as a function of altitude on Titan for frequencies of 100
Hz, 1000 Hz, 10000 Hz, and 1000000 Hz
absorption of 0.2 [5].
136
Figure 8.14. The scaled absorption for a 70 MHz signal on Earth (blue), Mars (red),
and Titan (green) at an altitude of 25 km compared to Navier-Stokes predicted thermal-
viscous losses (black line)
Chapter 9
Conclusions
The Direct Simulation Monte Carlo (DSMC) method has been used to numerically
investigate the problem of sound propagation in a variety of systems. DSMC is
capable of directly simulating acoustic waves with a level of detail not possible
with continuum approaches. The microscopic details found within the algorithm
have allowed for greater insight into the structure of an acoustics wave for all
Kn. Monatomic gases, gases with internal energy, planetary environments, and
amplitude eects spanning a large range of Kn have all been modeled with the
same method. The basis of DSMC lies within kinetic theory that is discussed in
detail in Chapter (2).
A parallel, object oriented DSMC solver was developed for this thesis. The
DSMC algorithm, including the implementation for acoustic wave propagation, was
described in Chapter (3). Assumptions and errors inherent in the algorithm were
also discussed. The code was written in C++ and is designed using the benets of
an object oriented approach. The parallelization is achieved by performing parallel
ensemble averaging, using the Message Passing Interface (MPI) library for inter-
processor communications. Despite the eciency of the parallel algorithm, the
CPU time for these DSMC programs can still be fairly large, especially for Kn <
0.05 situations or for problems with small deviations from equilibrium (


0
).
These computing issues were discussed in Chapter 4.
Chapter (5) derived and discussed many current theories for the absorption
and dispersion of sound in a simple monatomic gas as a function of Kn. The
results from several DSMC simulations in argon were compared and contrasted to
138
continuum theory. Results indicate that the absorption and dispersion of sound
are dependent on Kn and deviate signicantly from continuum theory at high Kn
but agree well with experimental values. The details of translational relaxation
were described in a monatomic gas for a range of Kn. The validity of the DSMC
method for all Kn allows for the study of nonequilibrium eects, even within the
continuum assumption. The ability to investigate the measure of nonequilibrium
in a system is unique to particle methods and that makes DSMC the method best
suited to investigate nonequilibrium eects.
Chapter (6) included theory on the losses associated with the redistribution
of rotational and vibrational energy. Several DSMC simulations in nitrogen with
only rotational energy were performed for a variety of relaxation collision numbers.
DSMC results follow relaxation theory within the continuum assumption and ap-
proach the free molecular limit for Kn > 1. The rotational relaxation peak is
evident in cases with higher relaxation collision number, but in all the other cases,
rotational relaxation occurs simultaneously with translational relaxation. In these
cases, the scaled absorption is more dependent on Kn indicating that there is a
unique combination of rotational relaxation and translational relaxation.
Chapter (6) also presents DSMC results comparing internal energy models as a
function of temperature in gaseous nitrogen with rotational and vibrational energy.
Large deviations from continuum theory were seen for all temperatures for Kn >
0.05, as expected. Dierences between internal energy modes were small for Kn <
0.05 at 0

C. The quantum and vibrationless congurations yield similar results for


all Kn at 0

C while the classical vibrational results are lower than the quantum
and vibrationless cases for Kn > 0.05. At 0

C, the quantum model for vibrational


energy is the most realistic, and should be the model of choice. At 2000 K and 4000
K, the dierences between the classical and quantum vibration models are small,
implying that the classical model can be a good approximation for vibrational
energy exchange at high temperatures.
Amplitude eects were discussed in Chapter (7). An investigation on the am-
plitude dependence of the scaled absorption coecient was performed for a wide
range of Kn and amplitudes. Again, large deviations from continuum theory is
seen for Kn > 0.05 as expected. Dierences in the absorption results between
amplitudes were small except for around the translational relaxation frequency at
139
a nondimensional distance of x/u

= 10, despite nonlinear eects that are evident


at low Kn. For distances beyond x/u

= 10 nonlinear and nonequilibrium eects


create nonexponential decay for high Kn, high amplitude sound. Nonequilibrium
eects were evident in the simulations, even at low Kn. The dispersion of sound
depended more uniquely on amplitude, depending on whether the data was pro-
cessed using the maximum acoustic pressure or the zero crossing of the acoustic
pressure. Several other nonlinear acoustics phenomenon such as harmonic genera-
tion, wave steepening and shock coalescence were also presented for a wide range
of Kn.
Investigations on three dierent celestial bodies, Earth, Mars and Titan, were
performed to describe various sound propagation properties for a wide range of Kn
in Chapter (8). DSMC results for a simple molecular relaxation were performed in
Earth-like conditions. Relaxation peaks for multiple relaxation collision numbers
were seen. For frequencies above the vibrational relaxation frequency, results on
Earth agreed very well with classical thermal-viscous theory. On Mars, DSMC
results show that the absorption of sound is 100 time greater on Mars than on
Earth and DSMC results agreed well with theory. Nonequilibrium eects are seen
for large Kn on Mars. Titans thick atmosphere is more conducive to making and
sustaining nonlinear waves. The Goldberg number on Titan is larger than on
Earth implying the importance of nonlinearity. The Knudsen number is large in
upper atmospheric conditions on all three planets and requires a particle method
solution due to the breakdown of the continuum assumption. DSMC should be the
computational method of choice for describing the acoustic phenomenon in high
altitude situations.
The conclusions of this work are summarized below:
1. Large discrepancies between experiment and theory still exist on the behav-
ior of sound at high Kn. The DSMC algorithm is the method best suited to
model noncontinuum sound propagation without relying on linearized (con-
tinuum or kinetic) theory.
2. This is the rst documented study to show the usefulness of DSMC over tra-
ditional continuum methods to study nonequilibrium and relaxation eects
on sound in gases with internal energy for the entire Kn range.
140
3. DSMC is a physically realistic model that includes nonlinearity, absorption,
relaxation, and internal energy eects which has been used to model sound
propagation in a variety of planetary environments where experimentation is
dicult. The use of DSMC to study these environments is the method best
suited to study high altitude eects.
4. Nonlinear eects beyond the continuum regime are poorly understood. The
DSMC method here is the only known study to model high amplitude non-
continuum sound.
5. A parallel, object-oriented, DSMC solver has been developed for acoustics
simulations. The solver has been used in a variety of systems including:
monatomic gases, gases with internal (quantum or classical) energy, and gas
mixtures, with little modication. The method could easily be implemented
in various other uid-ow problems as well.
6. Computation and memory costs are high, in particular for low Kn simula-
tions. Because of computer limitations, DSMC is currently limited to situ-
ations where Kn 0.0008. This implies that DSMC is currently not able
to simulate audible frequencies or large scale simulations on Earth. In addi-
tion, the vibrational relaxation collision numbers used in this study are not
physically realistic but where chosen based on the size of the system which
is a computational restriction.
9.1 Future work
The present study has explored the use of DSMC for acoustic wave propagation.
The results obtained from the simulations are promising. However, a lot of im-
provements can be made to extend the DSMC methods applications and to make
it more useful. The key areas that can be the focus of future research are identied
here.
As the scope of the research presented in this thesis was restricted to acoustic
waves generated in a one-dimensional simulation domain by a piston boundary
condition, there are many physical eects and phenomena which have been ex-
cluded from consideration. The extension to multiple dimensions would be fairly
141
trivial but would quickly become a memory intensive program. Therefore, the
parallelization implemented here would be prohibitive for multi-dimensional ows.
A domain decomposition algorithm would be better suited for this task. This type
of decomposition would also allow for the simulation of smaller Knudsen num-
bers. To circumvent the load balancing concern, the possibility of an adaptive grid
algorithm could also be implemented.
The piston-like boundary condition used in this study is only an approximation
to a piston face. The piston was simulated as a rigid wall where particle collisions
with the piston face would result in sinusoidally oscillating velocity components.
Alternatively, a moving oscillating solid wall can be used as the acoustic source.
Such a wall in conjunction with DSMC collision models that model non-hard-sphere
interactions is expected to be a better model of a typical experimental setup.
The extension of the model to include chemical reactions, ionized gases, or other
interesting nonhydrodynamic eects in binary mixtures (e.g., fast sound[165, 166])
would also be of interest.
Because of the easily adaptable DSMC algorithm, many other acoustical phe-
nomenon including but not limited to: mixture separation, boundary layer prob-
lems, streaming, boundary interactions, etc., could also be modeled. In addition,
other specic gas mixtures could be explored. By including multiple quantum vi-
brational levels or vibrational modes, a more realistic picture of relaxation due to
vibration could be understood.
With continued increase in computational power in addition to improvements to
the eciency of the algorithm (in terms of memory and calculations required), it is
very possible that the DSMC method will be capable of accurately modeling these
more complex multi-dimensional systems to aid in understanding the complexities
of sound propagation for all Knudsen numbers.
Appendix A
Buons Needle Experiment
Buons needle experiment is a simple Monte Carlo method for estimating the
value of . A grid of equally space parallel lines a distance d apart are drawn on a
table. A needle of length l d is dropped randomly on the table. The probability
that the needle would intersect a line is calculated by randomly throwing the needle
many times and calculating the ratio of the number of throws intersecting a line to
the total number of throws. Georges-Louis Leclerc, Comte de Buon (1707-1788),
a French naturalist and mathematician, solved this problem in 1777 [41].
Let x be the distance from the center of the needle to the closest line. Let be
the acute angle between the needle and the lines as shown in Fig. (A.1).
Figure A.1. An illustration of the Buons needle experiment.
143
The probability density function of x between 0 and d/2 is given by:
2
d
dx , (A.1)
and the probability density function of between 0 and /2 is:
2

d . (A.2)
The two random variables in this problem, x and , are independent and so
the joint probability density function is the product of Eqs. (A.1) and (A.2):
4
d
dxd . (A.3)
The needle will cross a line if:
x
l
2
sin . (A.4)
Integrating Eq. (A.3) will give the probability that the needle will cross a line.

2
_
0
(l/2) sin
_
0
4
d
dxd =
2l
d
. (A.5)
For n needles dropped with h of the needles crossing lines, the probability is given
by:
h
n
=
2l
d
(A.6)
which can be solved for .
The accuracy of the solution increases with the number of throws, and becomes
exact as the number of throws goes to innity.
Appendix B
Sampling a Maxwellian Distribution
The distribution of a random variable x may be described by a normalized dis-
tribution function such that the probability of a value of x lying between x and
x + dx is given by:
F
x
dx . (B.1)
If the range of x is from a to b the total probability is:
b
_
a
F
x
dx = 1 . (B.2)
The cumulative distribution function is dened as:
T
x
=
x
_
a
F
x
dx . (B.3)
Now a random fraction R
F
can now be generated and set equal to T
x
. Therefore,
T
x
= R
F
. (B.4)
In the trivial case that the variable x is uniformly distributed between a and b
then F
x
is a constant so that:
F
x
= 1/(b a) . (B.5)
145
Therefore, from Eq. (B.3),
T
x
=
b
_
a
1/(b a) = x a/(b a) (B.6)
and Eq. (B.4) gives:
(x a)/(b a) = R
F
(B.7)
which can be rewritten:
x = a + R
F
(b a) . (B.8)
Now consider the Maxwellian velocity distribution function, f
0
for the thermal
velocity in an equilibrium given by Eq. (2.58). The fraction of molecules with the
thermal velocity component in the x direction is given by integrating Eq. (2.58)
which gives the distribution function for the thermal velocity component u

x
as:
f
0
u

x
= n

exp
_

2
m
(u

x
)
2
)
_
, (B.9)
Consider the normalized Maxwellian distribution for the thermal velocity com-
ponent u

x
, F
0
u

x
so that:

F
0
u

x
du

x
= 1 . (B.10)
Therefore,
f
0
u

x
= nF
0
u

x
. (B.11)
Therefore, the normalized Maxwellian distribution function for the thermal
velocity component is written:
F
0
u

x
=

m

exp
_

2
m
(u

x
)
2
)
_
, (B.12)
A description of a direct method for sampling pairs of values, denoted u

x
and
v

x
, from the normal distribution of Eq. (B.12) will no be given. Let:
u

x
= r cos
v

x
= r sin . (B.13)
146
Since u

x
and v

x
are independent, the probability of u

x
lying in the range u

x
and u

x
+ du

x
and v

x
lying in the range v

x
and v

x
+ dv

x
is given by:
F
0
u

x
F
0
v

x
=

m

exp
_

2
m
(u

x
)
2
)
_
du

exp
_

2
m
(v

x
)
2
)
_
dv

x
=

2
m

exp
_

2
m
_
(u

x
)
2
+ (v

x
)
2
__
du

x
dv

x
. (B.14)
The Jacobian of the transformation is:
(u

x
, v

x
)
(r, )
=

x
r
u

x
r
v

cos r sin
sin r cos

= r . (B.15)
Therefore, we can write Eq. (B.14) as:
F
0
u

x
F
0
v

x
=

2
m

exp
_

2
m
r
2
_
rdrd . (B.16)
is a random variable uniformly distributed between 0 and 2 so that from
Eq. (B.8):
= 2R
F
. (B.17)
The variable
2
m
r
2
is distributed between 0 and and its distribution function:
F

2
m
r
2 = exp
_

2
m
r
2
_
, (B.18)
is already in normalized form. The cumulative distribution function for
2
m
r
2
can
be written:
T

2
m
r
2 = 1 exp
_

2
m
r
2
_
. (B.19)
Since R
F
and 1 R
F
are equivalent functions, Eq. (B.4) gives:
r = (ln (R
F
))
1/2
/
m
. (B.20)
A pair of values of r and may be sampled from Eqs. (B.17) and (B.20) using
successive random fractions. The normally distributed values of u

x
and v

x
follow
from Eq. (B.13) and provides a typical random value for the thermal velocity in
an equilibrium gas [1].
Appendix C
Deriving the Navier-Stokes
Dispersion Relation
The linearized conservation equations, Eqs. (5.13), (5.15), (5.17), and (5.19), can
be combined to create a secular equation with respect to the classical propagation
constant k
cl
. To see this, begin by substituting Eq. (5.13) into Eq. (5.15 ) to
obtain:
i(

c
2
0
p


0
T
0
T

) + ik
cl

0
U

x
= i

c
2
0
p

+ i

0
T
0
T

+ ik
cl

0
U

x
= 0 . (C.1)
Next, rewrite Eq. (5.17) and solve for p

to obtain:
p

= i
4
3
k
cl
U

x
+

0
U

x
k
cl
. (C.2)
Plugging Eq. (C.2) into Eq(C.1) yields:
i

c
2
0
_
i
4
3
k
cl
U

x
+

0
U

x
k
cl
_
+ i

0
T
0
T

+ ik
cl

0
U

x
= 0 . (C.3)
Rewrite Eq (5.19) and solve for T

.
T

=
p
0
ik
cl
U

x
i
0
c
v
k
2
cl

. (C.4)
148
Plug Eq. (C.4) into Eq (C.3) gives:
i

c
2
0
_
i
4
3
k
cl
U

x
+

0
U

x
k
cl
_
+ i

0
T
0
_
p
0
ik
cl
U

x
i
0
c
v
k
2
cl

_
+ ik
cl

0
U

x
= 0 . (C.5)
U

x
,= 0 so U

x
can be eliminated from the equation.
i

c
2
0
_
i
4
3
k
cl
+

0
k
_
+ i

0
T
0
_
p
0
ik
cl
i
0
c
v
k
2
cl

_
+ ik
cl

0
= 0 . (C.6)
Multiple both sides of the equation by i
0
c
v
k
2
cl
giving:
4k
cl
3c
2
0
(i
0
c
v
k
2
cl
)
i
2

0
c
2
0
k
cl
(i
0
c
v
k
2
cl
) +

0
p
0
k
cl
T
0
+ ik
cl

0
(i
0
c
v
k
2
) = 0 . (C.7)
Multiply both sides of the equation by k
cl
and expand to obtain:
i4
2

0
c
v
3c
2
0
k
2
cl

4
3c
2
0
k
4
cl

2
0
c
v
c
2
0
+
i
2

c
2
0
k
2
cl
+

0
p
0
T
0
k
2
cl
+
2
0
c
v
k
2
cl
i
0
k
4
cl
= 0 . (C.8)
In order to simplify Eq. (C.8) divide both sides by
2
0
c
v
.
i4
3c
2
0

0
k
2
cl

4
3c
2
0

2
0
c
v
k
4
cl

2
c
2
0
+
i
c
2
0

0
c
v
k
2
cl
+
p
0
T
0

0
c
v
k
2
cl
+
k
2
cl

0
c
v

k
4
cl
= 0 . (C.9)
Collect and order equal powers of k
cl
yields:

2
c
2
0
+
_
i4
3c
2
0

0
+
i
c
2
0

0
c
v
+
p
0
T
0

0
c
v
+
1

_
k
2
cl
+
_
4
3c
2
0

2
0
c
v

0
c
v

_
k
4
cl
= 0 . (C.10)
Some factoring gives:

2
c
2
0
+
_
i
c
2
0

0
_
4
3
+

c
v
_
+
p
0
T
0

0
c
v
+
1

_
k
2
cl
+

0
c
v
_
4
3c
2
0

_
k
4
cl
= 0 . (C.11)
Using the perfect gas law and thermodynamic properties,
p
0
T
0

0
cv
+
1

= 1 so
Eq. (C.11) becomes the dispersion relation given by the Navier-Stokes equations
149
for a simple gas written as a function of the propagation constant k
cl
as:


2
c
2
0
+
_
i
c
2
0

0
_
4
3
+

c
v
_
+ 1
_
k
2
cl
+

0
c
v
_
4
3c
2
0

_
k
4
cl
= 0 . (C.12)
Bibliography
[1] Bird, G. A. (1994) Molecular Gas Dynamics and the Direct Simulation of
Gas Flows, Clarendon Press, Oxford.
[2] Hanford, A. D., P. D. OConnor, J. B. Anderson, and L. N. Long
(2008) Predicting absorption and dispersion in acoustics by direct simula-
tion Monte Carlo: Quantum and classical models for molecular relaxation,
The Journal of the Acoustical Society of America, 123(6), pp. 41184126.
[3] Herzfeld, K. F. and T. A. Litovitz (1959) Absorption and Dispersion
of Ultrasonic Waves, Academic Press Inc., New York.
[4] Greenspan, M. (1956) Propagation of Sound in Five Monatomic Gases,
The Journal of the Acoustical Society of America, 28(4), pp. 644648.
[5] Schotter, R. (1974) Rareed gas acoustics in the noble gases, Physics
of Fluids, 17(6), pp. 11631168.
[6] Greenspan, M. (1954) Combined Translational and Relaxational Disper-
sion of Sound in Gases, The Journal of the Acoustical Society of America,
26(1), pp. 7073.
[7] Millikan, R. C. and D. R. White (1963) Systematics of Vibrational
Relaxation, The Journal of Chemical Physics, 39(12), pp. 32093213.
[8] Millikan, R. C. and D. R. White (1963) Vibrational Energy Exchange
between N
2
and CO. The Vibrational Relaxation of Nitrogen, The Journal
of Chemical Physics, 39(1), pp. 98101.
[9] Pierce, A. D. (1989) Acoustics - An Introduction to Its Physical Principles
and Applications, The Acoustical Society of America, Woodbury, New York.
[10] Hamilton, M. F. and D. T. Blackstock (eds.) (1998) Nonlinear Acous-
tics, Academic Press, San Diego.
151
[11] Online database for Earth available at http://www.spenvis.oma.be/ last
visited April, 2008.
[12] Online database for Mars available at http://www-mars.lmd.jussieu.fr/
last visited April, 2008.
[13] Bass, H. E. and J. P. Chambers (2001) Absorption of sound in the
Martian atmosphere, The Journal of the Acoustical Society of America,
109(6), pp. 30693071.
[14] Online database for Titan available at http://www.lmd.jussieu.fr/
titanDbase/ last visited April, 2008.
[15] Chapman, S. and T. G. Cowling (1970) The Mathematical Theory of
Non-Uniform Gases, third ed., Cambridge University Press, Cambridge.
[16] Cleveland, R. O., M. F. Hamilton, and D. T. Blackstock (1996)
Time-domain modeling of nite-amplitude sound in relaxing uids, The
Journal of the Acoustical Society of America, 99(6), pp. 33123318.
[17] Cleveland, R. O., J. P. Chambers, H. E. Bass, R. Raspet, D. T.
Blackstock, and M. F. Hamilton (1996) Comparison of computer
codes for the propagation of sonic boom waveforms through isothermal at-
mospheres, The Journal of the Acoustical Society of America, 100(5), pp.
30173027.
[18] Sparrow, V. W. and R. Raspet (1991) A numerical method for general
nite amplitude wave propagation in two dimensions and its application to
spark pulses, The Journal of the Acoustical Society of America, 90(5), pp.
26832691.
[19] Morris, P. J., L. N. Long, T. E. Scheidegger, and S. Boluriaan
(2002) Simulations of supersonic jet noise, International Journal of Aeroa-
coustics, 1, pp. 1741.
[20] Tam, C. K. W. and K. A. Kurbatskii (2003) Multi size meshmulti time
step dispersion relation preserving sheme for multiple scales aeroacoustics
problems, International Journal of Computational Fluid Dynamics, 17, pp.
119132.
[21] Tam, C. K. W. (2004) Computational aeroacoustics: An overview of com-
putational challenges and applications, International Journal of Computa-
tional Fluid Dynamics, 18, pp. 547567.
[22] Long, L. N., P. J. Morris, and A. Agarwal (2004) A review of par-
allel computing in computational aeroacoustics, International Journal of
Computational Fluid Dynamics, 18, pp. 493502.
152
[23] Kurbatskii, K. A. and R. R. Mankbadi (2004) Review of compu-
tational aeroacoustics algorithms, International Journal of Computational
Fluid Dynamics, 18, pp. 533546.
[24] Wochner, M. S., A. A. Atchley, and V. W. Sparrow (2005) Numer-
ical simulation of nite amplitude wave propagation in air using a realistic
atmospheric absorption model, The Journal of the Acoustical Society of
America, 118(5), pp. 28912898.
[25] Lighthill, M. J. (1952) On sound generated aerodynamically. I. General
theory, Proceedings of the Royal Society of London. Series A, Mathematical
and Physical Sciences, 211, pp. 564587.
[26] Lighthill , M. J. (1954) On sound generated aerodynamically. II. Tur-
bulence as a source of sound, Proceedings of the Royal Society of London.
Series A, Mathematical and Physical Sciences, 222, pp. 132.
[27] Hadjiconstantinou, N. G. (2006) The limits of Navier-Stokes theory
and kinetic extensions for describing small-scale gaseous hydrodynamics,
Physics of Fluids, 18(11), 111301.
[28] Frenkel, D. and B. Smit (1996) Understanding molecular simulation:
From algorithms to applications, Academic Press, San Diego.
[29] Allen, M. P. and D. J. Tilldesley (1987) Computer simulation of uids,
Oxford University Press, New York.
[30] Rapaport, D. C. (1995) The art of molecular dynamics simulation, Cam-
bridge University Press, Cambridge.
[31] Metropolis, N. and S. Ulam (1949) The Monte Carlo method, Journal
of the American Statistical Association, 44(247), pp. 335341.
[32] Aspray, W. (1990) John con Neumann and the origins of modern comput-
ing, MIT Press, Cambridge.
[33] Long, L. N., R. E. Coopersmith, and B. G. McLachlan (1987) Cel-
lular automatons applied to gas dynamic problems, in AIAA 19th Fluid
Dynamics, Plasma Dynamics, and Lasers Conference, American Institute of
Aeronautics and Astronautics, pp. AIAA paper 871384.
[34] Broadwell, J. E. (1964) Study of rareed shear ow by the discrete
velocity method, Journal of Fluid Mechanics, 19(03), pp. 401414.
[35] Broadwell , J. E. (1964) Shock Structure in a Simple Discrete Velocity
Gas, Physics of Fluids, 7(8), pp. 12431247.
153
[36] Chen, S. and G. D. Doolen (1998) Lattice Boltzmann method for uid
ows, Annual Review of Fluid Mechanics, 30(1), pp. 329364.
[37] Benzi, R., S. Succi, and M. Vergassola (1992) The lattice Boltzmann
equation: theory and applications, Physics Reports, 222, p. 146.
[38] J M Buick, C. A. G., C L Buckley and J. Gilbert (2000) Lattice
Boltzmann BGK simulation of nonlinear sound waves: the development of a
shock front, Journal of Physics A: Mathematical and General, 33(21), pp.
39173928.
[39] Sudo, Y. and V. W. Sparrow (1993) A new lattice gas model for 1-D
sound propagation, Journal of Computational Acoustics, 1(4), pp. 423454.
[40] Sudo, Y. and V. W. Sparrow (1995) Lattice gas wave propagation
models including dissipation, Journal of Computational Acoustics, 3(1), pp.
6993.
[41] de Buffon, G. C. (1777) Essi darithmetique morale, Supplement `a
lHistorie Naturelle, 4.
[42] Kelvin, L. (1901) Nineteenth century clouds over the dynamical theory of
heat and light, Philosophical Magazine, 2(6), pp. 140.
[43] Bird, G. A. (1963) Approach to translational equilibrium in a rigid sphere
gas, Physics of Fluids, 6, pp. 15181519.
[44] Chang, C. S. W. and G. E. Uhlenbeck (1970) Studies in Statistical
Mechanics, vol. V, North Holland, Amsterdam, pp. 4375.
[45] Pekeris, C. L., Z. Alterman, L. Finkelstein, and K. Frankowski
(1962) Propagation of Sound in a Gas of Rigid Spheres, Physics of Fluids,
5(12), pp. 16081616.
[46] Sirovich, L. and J. K. Thurber (1965) Propagation of Forced Sound
Waves in Rareed Gasdynamics, The Journal of the Acoustical Society of
America, 37(2), pp. 329339.
[47] Loyalka, S. K. and T. C. Cheng (1979) Sound-wave propagation in a
rareed gas, Physics of Fluids, 22(5), pp. 830836.
[48] Marques, W. (1999) Dispersion and absorption of sound in monatomic
gases: An extended kinetic description, The Journal of the Acoustical So-
ciety of America, 106(6), pp. 32823288.
154
[49] Maidanik, G., H. L. Fox, and M. Heckl (1965) Propagation and Re-
ection of Sound in Rareed Gases. I. Theoretical, Physics of Fluids, 8(2),
pp. 259265.
[50] Buckner, J. K. and J. H. Ferziger (1966) Linearized Boundary Value
Problem for a Gas and Sound Propagation, Physics of Fluids, 9(12), pp.
23152322.
[51] Zuckerwar, A. J. (1999) Phenomenological theory of the translational
relaxation times in gases, The Journal of the Acoustical Society of America,
105(4), pp. 22102215.
[52] Greenspan , M. (1959) Rotational Relaxation in Nitrogen, Oxygen, and
Air, The Journal of the Acoustical Society of America, 31(2), pp. 155160.
[53] Meyer, E. and G. Sessler (1957) Schallausbreitung in Gasen bei hohen
Frequenzen und sehr niedrigen Drucken, Zeitschrift f ur Physik A: Hadrons
und Nuclei, 149(1), pp. 1539.
[54] Maidanik, G. and M. Heckl (1965) Propagation and Reection of Sound
in Rareed Gases. II. Experimental, Physics of Fluids, 8(2), pp. 266272.
[55] Herzfeld, K. F. and F. O. Rice (1928) Dispersion and Absorption of
High Frequency Sound Waves, Phys. Rev., 31(4), pp. 691695.
[56] Kneser, H. O. (1933) The Interpretation of the Anomalous Sound-
Absorption in Air and Oxygen in Terms of Molecular Collisions, The Jour-
nal of the Acoustical Society of America, 5(2), pp. 122126.
[57] Chang, C. S. W. and G. E. Uhlenbeck (1951) Transport Phenomena
in Polyatomic Gases, Tech. Rep. Research Report No. CM-681, University
of Michigan Engineering.
[58] Monchick, L., K. S. Yun, and E. A. Mason (1963) Formal Kinetic
Theory of Transport Phenomena in Polyatomic Gas Mixtures, The Journal
of Chemical Physics, 39(3), pp. 654669.
[59] Mason, E. A. and L. Monchick (1962) Heat Conductivity of Polyatomic
and Polar Gases, The Journal of Chemical Physics, 36(6), pp. 16221639.
[60] Hanson, F. B. and T. F. Morse (1967) Kinetic Models for a Gas with
Internal Structure, Physics of Fluids, 10(2), pp. 345353.
[61] Hanson, F. B., T. F. Morse, and L. Sirovich (1969) Kinetic Descrip-
tion of the Propagation of Plane Sound Waves in a Diatomic Gas, Physics
of Fluids, 12(1), pp. 8495.
155
[62] McCormack, F. J. and H. W. Creech (1972) 17-Moment Theory of
Sound Propagation in Polyatomic Gases, The Journal of the Acoustical
Society of America, 51(3B), pp. 900909.
[63] McCormack, F. J. (1972) Greenspan Rule for Combined Absorption of
Sound in Polyatomic Gases with Many Internal Degrees of Freedom, The
Journal of the Acoustical Society of America, 52(5B), pp. 13771379.
[64] Banankhah, A. and S. K. Loyalka (1987) Propagation of a sound wave
in a rareed polyatomic gas, Physics of Fluids, 30(1), pp. 5664.
[65] Danforth, A. L. and L. N. Long (2004) Nonlinear acoustic simulations
using direct simulation Monte Carlo, The Journal of the Acoustical Society
of America, 116(4), pp. 19481955.
[66] Alexander, F. J. and A. L. Garcia (1997) The direct simulation Monte
Carlo method, Computers in Physics, 11, p. 588.
[67] Wagner, W. (1992) A convergence proof for Birds direct simulation
Monte Carlo method for the Boltzmann equation, Journal of Statistical
Physics, 66(3-4), pp. 10111044.
[68] Muntz, E. P. (1989) Rareed Gas Dynamics, Annual Review of Fluid
Mechanics, 21(1), pp. 387422.
[69] Bird, G. A. (1978) Monte Carlo Simulation of Gas Flows, Annual Review
of Fluid Mechanics, 10(1), pp. 1131.
[70] Oran, E., C. Oh, and B. Cybyk (1998) Direct Simulation Monte Carlo:
Recent Advances and Applications, Annual Review of Fluid Mechanics,
30(1), pp. 403441.
[71] Bird, G. A. (1970) Direct Simulation and the Boltzmann Equation,
Physics of Fluids, 13(11), pp. 26762681.
[72] Long, L. N. (1991) Navier Stokes and Monte Carlo results for hypersonic
ows, AIAA Journal, 29(2), pp. 200207.
[73] Hanford, A. D. and L. N. Long (hopefully 2008ish) The direct simula-
tion of acoustics on Earth, Mars, and Titan, The Journal of the Acoustical
Society of America.
[74] Hadjiconstantinou, N. G. and A. L. Garcia (2001) Molecular simu-
lations of sound wave propagation in simple gases, Physics of Fluids, 13(4),
pp. 10401046.
156
[75] OConnor, P. D., L. N. Long, and J. B. Anderson (2008) Accurate
Rate Expressions for Simulations of Gas-Phase Chemical Reactions, Journal
of Computational Physics, 227, pp. 76647673.
[76] Ikegawa, M. and J. Kobayashi (1989) Deposition prole simulation us-
ing the direct simulation Monte Carlo method, Journal of the Electrochem-
ical Society, 136, pp. 303310.
[77] Coronell, D. G. and K. F. Jensen (1989) Monte Carlo simulations of
very low pressure chemical vapor desposition, Journal of Computer-Aided
Material Design, 1, pp. 326.
[78] OConnor, P. D., L. N. Long, and J. B. Anderson (2006) The Direct
Simulation of Detonations, in AIAA/ASME/SAE/ASEE Joint Propulsion
Conference, American Institute of Aeronautics and Astronautics, pp. AIAA
Paper No. 20064411.
[79] Anderson, J. B. and L. N. Long (2003) Direct Monte Carlo simula-
tion of chemical reaction systems: Prediction of ultrafast detonations, The
Journal of Chemical Physics, 118(7), pp. 31023110.
[80] Zhang, J., D. B. Goldstein, P. L. Varghese, L. Trafton,
C. Moore, and K. Miki (2004) Numerical modeling of ionian volcanic
plumes with entrained particulates, Icarus, 172, pp. 479502.
[81] Austin, J. V. and D. B. Goldstein (1997) Direct numerical simulation
of circumplanetary winds on Io, Bulletin of the American Astronomical
Society, 29, p. 1004.
[82] Yang, D. (2001) C++ and Object-Oriented Numeric Computing, Springer,
New York.
[83] Shtern, V. (2000) Core C++, Prentice Hall PTR, Upper Saddle River.
[84] Pacheco, P. S. (1997) Parallel Programming with MPI, Morgan Kaufm-
mann Publishers, San Fransisco.
[85] Gropp, W., E. Lusk, and A. Skjellum (1998) Using MPI: Portable Par-
allel Programming with Message Passing Interface, MIT Press, Cambridge.
[86] Danforth-Hanford, A., P. D. OConnor, L. N. Long, and J. B. An-
derson (2006) Molecular Relaxation Simulations in Nonlinear Acoustics
using Direct Simulation Monte Carlo, in INNOVATIONS IN NONLINEAR
ACOUSTICS: ISNA17 - 17th International Symposium on Nonlinear Acous-
tics including the International Sonic Boom Forum (A. A. Atchley, V. W.
Sparrow, and R. M. Keolian, eds.), vol. 838, American Institute of Physics,
pp. 556559.
157
[87] Hanford, A. D., L. N. Long, and V. W. Sparrow (2007) The prop-
agation of sound on Titan using the direct simulation Monte Carlo, The
Journal of the Acoustical Society of America, 121(5), p. 3117.
[88] Hanford, A. D. and L. N. Long (2006) The absorption of sound on
Mars using the direct simulation Monte Carlo, The Journal of the Acoustical
Society of America, 119(5), p. 3264.
[89] Lindley, D. (2001) Boltzmanns atom: the great debate that launched a
revolution in physics, Free Press.
[90] Brush, S. G. (2003) The kinetic theroy of gases: an anthology of classic
papers with historical commentary, Imperial College Press.
[91] Perrin, J. B. (1909) Le Mouvement Brownien et la Realite Monleculaire,
Annales de Chimie Physique, 18, pp. 5114.
[92] Einstein, A. (1905)

Uber die von der molekularkinetischen Theorie der


Waarme geforderte Bewegung von in ruhenden Fl ussigkeiten suspendierten
Teilchen, Annalen der Physik, 322(8), pp. 549560.
[93] Anderson, J. D. (1998) A history of aerodynamics and its impact on ying
machines, Cambridge University Press.
[94] Harris, S. (1971) An introduction to the theory of the Boltzmann equation,
Holt, Rinehart, and Winston, New York.
[95] Long, L. N., M. Kamon, and J. Myczkowski (1992) A massively par-
allel algorithm to solve the Boltzmann (BGK) equation, in AIAA Aerospace
Sciences Meeting, American Institute of Aeronautics and Astronautics, pp.
AIAA paper 920563.
[96] Long, L. N. and J. Myczkowski (1993) Solving the Boltzmann equation
at 61 gigaops on a 1024-node CM-5, in IEEE proceedings of supercomput-
ing.
[97] Chapman, S. (1918) On the Kinetic Theory of a Gas. Part II: A Compos-
ite Monatomic Gas: Diusion, Viscosity, and Thermal Conduction, Philo-
sophical Transactions of the Royal Society of London. Series A, Containing
Papers of a Mathematical or Physical Character, 217, pp. 115197.
[98] Enskog, D. (1921) Kinetische Theorie der Vorgange in massig verdunnten
Gasen, Arkiv. Mat. Astr. Fys., 16, pp. 160.
[99] Agarwal, R. K., K.-Y. Yun, and R. Balakrishnan (2001) Beyond
NavierStokes: Burnett equations for ows in the continuumtransition
regime, Physics of Fluids, 13(10), pp. 30613085.
158
[100] Spiegel, E. A. and J.-L. Thiffeault (2003) Higher-order continuum
approximation for rareed gases, Physics of Fluids, 15(11), pp. 35583567.
[101] Sonine, N. J. (1880) Recherches sur les fonctions cylindriques et le
developpement des fonctions continues en series, Mathematische Annalen,
16, pp. 180.
[102] Spiegel, M. R. and J. Liu (1999) Mathematical Handbook of Formulas
and Tables, 2nd edition ed., McGraw-Hill, New York.
[103] Burnett, D. (1935) The distribution of velocities in a slightly non-uniform
gas, Proceedings of the London Mathematical Society, 39, pp. 385430.
[104] Dorrance, W. H. (1962) Viscous Hypersonic Flow, McGraww-Hill.
[105] Bhatnagar, P. L., E. P. Gross, and M. Krook (1954) A Model for
Collision Processes in Gases. I. Small Amplitude Processes in Charged and
Neutral One-Component Systems, Phys. Rev., 94(3), pp. 511525.
[106] Bellman, R. E. (2003) Dynamic Programming, reprint edition ed., Dover
Publications.
[107] Greenspan, M. and M. C. T. Jr (1953) An Eleven Megacycle Interfer-
ometer for Low Pressure Gases, The Journal of the Acoustical Society of
America, 25(1), pp. 9296.
[108] Sharipov, F., J. W. Marques, and G. M. Kremer (2002) Free molec-
ular sound propagation, The Journal of the Acoustical Society of America,
112(2), pp. 395401.
[109] Borgnakke, C. and P. S. Larsen (1975) Statistical collision model for
Monte Carlo simulation of polyatomic gas mixture, Journal of Computa-
tional Physics, 18(4), 405.
[110] Anderson, J. B., J. D. Foch, M. J. Shaw, R. C. Stern, and B. J.
Wu (1986) Statistical theory of electronic energy relaxation, in Proceedings
from the 15
th
International Symposium on Rareed Gas Dynamics (V. Bo
and C. Cercignani, eds.), vol. 2, B. G. Teubner Stuttgart, pp. 413421.
[111] Bergemann, F. and I. D. Boyd (1992) New Discrete Vibrational Energy
Model for the Direct Simulation Monte Carlo Method, in Proceedings from
the 18
th
International Symposium on Rareed Gas Dynamics (B. D. Shizgal
and D. P. Weaver, eds.), vol. 158, American Institute of Aeronautics and
Astronautics, pp. 174183.
[112] Garcia, A. L. and W. Wagner (2000) Time step truncation error in
direct simulation Monte Carlo, Physics of Fluids, 12(10), pp. 26212633.
159
[113] Hadjiconstantinou, N. G. (2000) Analysis of discretization in the direct
simulation Monte Carlo, Physics of Fluids, 12(10), pp. 26342638.
[114] Alexander, F. J., A. L. Garcia, and B. J. Alder (1998) Cell size de-
pendence of transport coecients in stochastic particle algorithms, Physics
of Fluids, 10(6), pp. 15401542.
[115] Alexander, F. J., A. L. Garcia, and B. J. Alder (2000) Erra-
tum: Cell size dependence of transport coecients in stochastic particle
algorithms [Phys. Fluids 10, 1540 (1998)], Physics of Fluids, 12(3), pp.
731731.
[116] Hadjiconstantinou, N. G., A. L. Garcia, M. Z. Bazant, and G. He
(2003) Statistical error in particle simulations of hydrodynamic phenom-
ena, Journal of Compuational Physics, 187, pp. 274297.
[117] Long, L. N. and J. B. Anderson (2001) The simulation of detonations
using a Monte Carlo method, vol. 585, American Institute of Physics, New
York, pp. 653657.
[118] Long, L. N. and Brentner (2000) Self-Scheduling parallel methods
for mulitple serial codes with application to WOPWOP, in AIAA 38th
Aerospace Sciences Meeting and Exhibit, American Institute of Aeronautics
and Astronautics, pp. AIAA 20000346.
[119] Hardware description and documentation available at: http://www.cse.
psu.edu/HPC/mufasa/ last visited April 2008.
[120] Hardware description and documentation available at: http://www.nas.
nasa.gov/Resources/Systems/columbia.html last visited April 2008.
[121] Top 500 supercomputers listed at: http://www.top500.org last visited
April 2008.
[122] Sharma, A. and L. N. Long (2004) Numerical simulation of the blast
impact problem using the direct simulation Monte Carlo (DSMC) method,
Journal of Computational Physics, 200, pp. 211237.
[123] Fan, J. and C. Shen (2001) Statistical simulation of low-speed rareed
gas ows, Journal of Computational Physics, 167, pp. 393412.
[124] Pan, L. S., T. Y. Ng, D. Xu, and K. Y. Lam (2001) Molecular
block model direct simulation Monte Carlo method for low velocity microgas
ows, Journal of Micromechanics and Microengineering, 11, pp. 181188.
[125] Sun, Q. and I. D. Boyd (2002) A direct simulation method for subsonic,
microscale gas ows, Journal of Computational Physics, 179, pp. 400425.
160
[126] Homolle, T. M. M. and N. G. Hadjiconstantinou (2007) Low-
variance deviational simulation Monte Carlo, Physics of Fluids, 19(4),
041701.
[127] Wang, M. and Z. Li (2004) Failure analysis of the molecular block model
for the direct simulation Monte Carlo method, Phyiscs of Fluids, 16, pp.
21222125.
[128] Baker, L. L. and N. G. Hadjiconstantinou (2005) Variance reduction
for Monte Carlo solutions of the Boltzmann equation, Physics of Fluids,
17(5), 051703.
[129] Pullin, D. I. (1980) Direct simulation methods for ideal-gas ow, Journal
of Computational Physics, 34, pp. 231244.
[130] Hadjiconstantinou, N. G. (2002) Sound wave propagation in transition-
regime micro- and nanochannels, Physics of Fluids, 14(2), pp. 802809.
[131] Sun, Q., I. D. Boyd, and G. V. Candler (2004) A hybrid contin-
uum/particle approach for modeling subsonic, rareed gas ows, Journal of
Computational Physics, 194, pp. 256277.
[132] Aktas, O. and N. R. Aluru (2002) A combined continuum/ DSMC
technique for multiscale analysis of microuidic lters, Journal of Compu-
tational Physics, 178, pp. 342372.
[133] Lenihan, J. M. A. (1951) Mersenne and Gassendi: an early chapter in
the theory of sound, Acustica, 2, pp. 9699.
[134] Hunt, F. V. (1987) Origins in Acoustics, The Acoustical Society of America
through the American Institute of Physics.
[135] Bass, H. E., L. C. Sutherland, J. Piercy, and L. Evans (1984) Physi-
cal Acoustics; Principles and Methods, vol. XVII, chap. Absorption of sound
by the atmosphere, Academic Press, Inc, pp. 145232.
[136] Duff, A. W. (1900) The Attenuation of Sound, Phys. Rev. (Series I),
11(2), pp. 6574.
[137] Knudsen, V. O. (1933) The Absorption of Sound in Air, in Oxygen, and
in NitrogenEects of Humidity and Temperature, The Journal of the
Acoustical Society of America, 5(2), pp. 112121.
[138] Greenspan, M. (1965) Physical Acoustics; Principles and Methods, vol. II,
chap. Transmission of Sound Waves in Gases at Very Low Pressures, Aca-
demic Press, Inc, pp. 143.
161
[139] Rayleigh, L. (1945) Theory of Sound, second ed., Dover Publications, New
York.
[140] Buckner, J. K. and J. H. Ferziger (1966) Linearized Initial Value
Problem for a Gas, Physics of Fluids, 9(12), pp. 23092314.
[141] Bauer, H. J. (1972) Inuences of transport mechanisms on sound propa-
gation in gases, Advances in Molecular Relaxation Processes, 2, pp. 319376.
[142] Markham, J. J., R. T. Beyer, and R. B. Lindsay (1951) Absorption
of Sound in Fluids, Rev. Mod. Phys., 23(4), pp. 353411.
[143] Bass, H. E. and L. C. Sutherland (1976) On the rotational collison
number for air at elevated temperatures, The Journal of the Acoustical
Society of America, 59(6), pp. 13171318.
[144] Sutherland, L. C. and H. E. Bass (2004) Atmospheric absorption in
the atmosphere up to 160 km, The Journal of the Acoustical Society of
America, 115(3), pp. 10121032.
[145] Bass, H. E., L. C. Sutherland, A. J. Zuckerwar, D. T. Black-
stock, and D. M. Hester (1995) Atmospheric absorption of sound: Fur-
ther developments, The Journal of the Acoustical Society of America, 97(1),
pp. 680683.
[146] Shields, F. D., C. C. Warf, and H. E. Bass (1973) Acoustical method
of obtaining vibrational transition rates tested on CO[sub 2]/N[sub 2] mix-
tures, The Journal of Chemical Physics, 58(9), pp. 38373840.
[147] Evans, L. B. (1972) Vibrational Relaxation in Moist Nitrogen, The Jour-
nal of the Acoustical Society of America, 51(1B), pp. 409411.
[148] Bass, H. E. and F. D. Shields (1974) Vibrational relaxation and sound
absorption in O
2
/H
2
O mixtures, The Journal of the Acoustical Society of
America, 56(3), pp. 856859.
[149] Winter, T. G. and G. L. Hill (1967) High-Temperature Ultrasonic Mea-
surements of Rotational Relaxation in Hydrogen, Deuterium, Nitrogen, and
Oxygen, The Journal of the Acoustical Society of America, 42(4), pp. 848
858.
[150] Blackstock, D. T. (1964) Thermoviscous Attenuation of plane, periodic,
nite-amplitude sound waves, The Acoustical Society of America, 36(3), pp.
534542.
[151] Mahaffy, P. R. (2005) Intensive Titan Exploration Begins. Science,
308(5724), pp. 969 970.
162
[152] Lebreton, J.-P., O. Witasse, C. Sollazzo, T. Blancquaert,
P. Couzin, A.-M. Schipper, J. B. Jones, D. L. Matson, L. I.
Gurvits, D. H. Atkinson, B. Kazeminejad, and M. P erez-Ay ucar
(2005) An overview of the descent and landing of the Huygens probe on
Titan, Nature, 438, pp. 758764.
[153] Zarnecki, J. C., M. R. Leese, J. R. C. Garry, N. Ghafoor, and
B. Hathi (2002) Huygens surface science package, Space Science Review,
104, pp. 593 611.
[154] Lorenz, R. D. (1998) Speed of sound in outer planet atmospheres, Plan-
etary and Space Science, 47(1-2), pp. 6777.
[155] Information on the Mars Microphone available at: http://sprg.ssl.
berkeley.edu/marsmic/welcome.html last visited April, 2008.
[156] Williams, J.-P. (2001) Acoustic enviornment of the Martian surface,
Journal of Geophysical Research, 106(E3), pp. 50335041.
[157] Petculescu, A. and R. M. Lueptow (2007) Atmospheric acoustics of
Titan, Mars, Venus, and Earth, Icarus, 186(2), pp. 413419.
[158] Forget, F., F. Hourdin, R. Fournier, C. Hourdin, and O. Tala-
grand (1999) Improved general circulation models of the Martian atmo-
sphere from the surface to above 80 km, Journal of Geophysical Research,
104, pp. 2415524175.
[159] Lewis, S. R., M. Collins, P. L. Read, F. Forget, F. Hourdin,
R. Fournier, C. Hourdin, O. Talagrand, and J.-P. Hout (1999)
A climate database for Mars, Journal of Geophysical Research, 104, pp.
2417724194.
[160] Rannou, P., S. Lebonnois, F. Hourdin, and D. Luz (2005) Titan
atmosphere database, Advances in Space Research, 36(11), pp. 21942198.
[161] Owen, T., K. Biemann, D. R. Rushneck, J. E. Biller, D. W.
Howarth, and A. L. Lafleur (1977) The composition of the atmosphere
at the surface of Mars, Journal of Geophysical Research, 82, pp. 46354639.
[162] Sparrow, V. W. (1999) Acoustics on the planet Mars: A preview, The
Journal of the Acoustical Society of America, 106(4), pp. 22642264.
[163] Flasar, F. M., R. K. Achterberg, B. J. Conrath, P. J. Gierasch,
V. G. Kunde, C. A. Nixon, G. L. Bjoraker, D. E. Jennings, P. N.
Romani, A. A. Simon-Miller, B. Bzard, A. Coustenis, P. G. J.
Irwin, N. A. Teanby, J. Brasunas, J. C. Pearl, M. E. Segura,
163
R. C. Carlson, A. Mamoutkine, and P. J. Schinder (2005) Titans
Atmospheric Temperatures, Winds, and Composition. Science, 308(5724),
pp. p975 978.
[164] Dain, Y. and R. M. Lueptow (2001) Acoustic attenuation in three-
component gas mixturesTheory, The Journal of the Acoustical Society of
America, 109(5), pp. 19551964.
[165] Huck, R. J. and E. A. Johnson (1980) Possibility of Double Sound
Propagation in Disparate-Mass Gas Mixtures, Phys. Rev. Lett., 44(3), pp.
142145.
[166] Montfrooij, W., P. Westerhuijs, V. O. de Haan, and I. M.
de Schepper (1989) Fast sound in a helium-neon mixture determined
by neutron scattering, Phys. Rev. Lett., 63(5), pp. 544546.
Vita
Amanda Danforth Hanford
Amanda Danforth was born on September 26, 1980 in Landstuhl, Germany.
With parents in the military, she moved a few times during childhood. Amanda
spent the majority of her young life in Colorado Springs, Colorado and graduated
with honors from Air Academy High School in 1998. Beginning in August 1998,
Amanda attended the University of Rochester in Rochester, New York where she
graduated Magne Cum Laude with a Bachelors of Arts in Mathematics and a
minor in Music in May 2002. She enrolled for graduate studies in the Graduate
Program in Acoustics at the Pennsylvania State University in August 2002. She
completed her Masters of Science in 2004 under the guidance of Prof. Lyle N. Long.
Her masters work was on Monte Carlo methods for nonlinear acoustics problems.
Amanda continued work using Monte Carlo methods with Prof. Lyle N. Long for
her doctoral studies beginning in 2004. Amanda also got married in August of
2004 and her married name became Amanda Danforth Hanford. Amanda had the
opportunity to teach 6 weeks of an undergraduate course in the fundamentals of
Acoustics in the spring of 2006. She also gave birth to a beautiful baby boy in
November of 2006. Amanda has a job oer as a research faculty at the Applied
Research Lab at the Pennsylvania State University.

You might also like