You are on page 1of 231

Extended Navier-Stokes Equations:

Derivations and Applications to Fluid


Flow Problems

Erweiterte Navier-Stokes-Gleichungen:
Ableitungen und Anwendungen auf
Strömungsprobleme

Der Technischen Fakultät der


Universität Erlangen-Nürnberg

zur Erlangung des Grades

DOKTOR-INGENIEUR

vorgelegt von

Rajamani Sambasivam

Erlangen, 2013

i
Als Dissertation genehmigt
von der Technischen Fakultät der
Universität Erlangen – Nürnberg

Tag der Einreichung : 14.09.2011

Tag der Promotion : 09.03.2012

Dekan : Prof. Dr.-Ing. habil. Marion Merklein

Berichterstatter : Prof. i. R. Dr. Dr. hc. Franz Durst


Prof. Dr. rer. nat. habil.Ulrich Rüde
PD. Dr.–Ing. habil. Stefan Becker

ii
Extended Navier-Stokes Equations:
Derivations and Applications to Fluid Flow
Problems

Erweiterte Navier-Stokes-Gleichungen:
Ableitungen und Anwendungen auf
Strömungsprobleme

PhD Thesis

Rajamani Sambasivam
ii
I dedicate this work

To the memory of my father, Mr. Rajamani


To my mother, Mrs. Gandhi
To my wife, Debjani
To my daughter, Aishani
To my son, Aadharsh
To my brother, Krishnan

iii
iv
ACKNOWLEDGEMENTS

This thesis is based on the work I carried out partly during my stay at the Institute of
Fluid Mechanics (LSTM), University of Erlangen, Germany during a collaborative
project and mostly during the brief stay at FMP Technology GmbH, Erlangen and at
Jamshedpur, India.

First of all, I would like to express my gratitude to my mentor and guide, Prof. Dr. Dr.
h. c. Franz Durst, for the guidance and support regarding this work and beyond. I am
also grateful to him for choosing me as a student and I will always cherish, throughout
my life, the interactions I had with him both in Germany and India. During the last
stages of my stay at LSTM, in the summer of 2005, he introduced me to this fascinating
world of extended Navier-Stokes equations in the sidelines of our meeting regarding
supersonic jet simulations expanding into hot environments. From that day, he
continued to support me in the pursuit of this exciting research domain and showed
child-like enthusiasm whenever I came up with some interesting results. Without his
constant encouragement and guidance, this work would not have reached this stage.
Further, I would also like to express my gratitude to him for providing insights into fluid
mechanics and transport processes which helped me to understand the underlying
physical phenomena ‘differently’.

The next person, who showed equal enthusiasm in my research, is Dr. Sanjay Chandra,
Chief of Technology (Global Wires and Longs), and my superior at TATA Steel. He
has been a constant source of encouragement and motivating force for me to pursue my
research interests in fundamental fluid mechanics along with applied research for TATA
Steel. I always value the support and guidance he has been providing all through my
tenure at TATA Steel. Further, I express my gratitude to all my teachers during my
schooling and also in graduation and post-graduation engineering studies. Especially, I
would like to name (i) Prof. K Subramanian, who taught me engineering
thermodynamics and thermal engineering at the undergraduate level at Government
College of Engineering, Tirunelveli, India and (ii) Prof. V. V. Satyamurty, Indian
Institute of Technology, Khargpur, who was my guide during my post-graduation thesis
on heat transfer. I thank them for the great insight they provided in two of my favourite
subjects.

I also would like to thank Dr. T Venugopalan, Chief Technology Officer, TATA Steel
for the constant encouragement and his frequent inquiries about my PhD. Further, I
would like to thank the management of TATA Steel for giving me an opportunity to
work in Germany for a collaborative project in the area of designing supersonic oxygen
lance for LD steelmaking and I might not have met Prof. Durst without this opportunity.
For this, I would like to thank Dr. T Mukherjee, erstwhile Deputy Managing Director,
TATA Steel, Dr. D. Bhattacharjee, then Chief of R&D and Scientific Services, Dr. S K
Ajmani, Head – Steelmaking and Casting research group and most importantly, once

v
again to Dr. Sanjay Chandra, then Chief of R&D. I also thank Mr. Sudhir Malavade of
TATA Steel who helped me with the corrections during submission of the thesis.

I would like to express my gratitude to Mrs. Heidi Durst, wife of Prof. Durst for
providing a homely environment for my family in Germany and also for accepting my
intrusions on Prof. Durst’s time on holidays and weekends. I also like to thank Ms.
Susanne Braun and Mrs. Johanna Grasser, FMP Technology for making all the long
written communications between me and Prof. Durst possible and also for sending
corrections and suggestions on the thesis from Prof. Durst available to me in time. I
would also like to thank Mr. Nishanth Dongari and Mr. C Karthik with whom I had
long discussions about the extended Navier-Stokes equations at FMP Technology
GmbH in Erlangen. I also express my thanks to Mr. Navaneetha Krishnan, Mr. D.
Filimonov and Dr. T. Adachi for making the analytical solutions to microchannel flows
possible and also for various technical discussions which helped me to understand the
microchannel flows better. I express my gratitude to Prof. Suman Chakraborty, Indian
Institute of Technology, Kharagpur, India for providing the DSMC simulations data for
the microchannel with a backward facing step.

I would like to thank another important person, who was one of my great motivators
during the progress of this work and that is Prof. Manoranjan Maiti, my father in-law. I
thank him for his constant encouragement, frequent inquiries about the status of the
work and suggesting corrections and changes during the writing of the thesis. Further, I
also thank Mr. M. B. Das who also provided a great degree of support and
encouragement during the course of this work.

Last but not the least, I express my gratitude and love to my whole family for being the
driving force in my life. I would like to thank my wife, Debjani for managing the house
and children during the frequent absence from home during various stages of this thesis
and also during business trips. I am grateful to her for keeping me focused in my work
during tough times with her support, suggestions and love. I thank my children, Aishani
and Aadharsh for providing the much needed diversion from the work with their love
and affection; my brother, Krishnan for his constant support and encouragement and to
my mother, Mrs. Gandhi Rajamani for her unwavering and impeccable support
throughout my life. Finally, I would like to dedicate this work and all my successes in
life to my late father, Mr. S. K. S. Rajamani who had been and will remain the greatest
father one could ever have. I thank him for his faith and confidence in me and for
urging me to do well in my studies and in whatever I venture into. He was the person
who I looked up to right from my childhood in my hours of need and now, his
memories have taken up that place.

14 September 2011
th
Rajamani Sambasivam

vi
ABSTRACT

The present thesis summarises the author’s research work, carried out under the
supervision of Prof. Franz Durst, Professor Emeritus, University of Erlangen-Nürnberg,
in the field of extended Navier-Stokes equations. Through some of the earlier
collaborative work, the author realised that under certain flow situations, the description
provided by the classical Navier-Stokes equations did not corroborate with the
corresponding experimental measurements. This had puzzled the author and led to a
closer look at the derivations of transport equations from the first principles. It was
found out that the derivations of the classical equations had certain flaws when
employed to solve gas flows with strong density/pressure and temperature gradients
which resulted in additional diffusion mass fluxes. It was argued that this additional
mass flux needed to be superimposed onto the convective mass fluxes treated in the
classical Navier-Stokes equations. Further, it was also necessary to incorporate
modified constitutive relationships of molecular transport of momentum and heat in the
governing equations. Hence it may be argued that the classical Navier-Stokes equations
are only valid when there are no density and temperature gradients in ideal gas flows.

The present thesis commences with a summary of development of the classical Navier-
Stokes equations based on the existing historical knowledge of fluid mechanics. A brief
history of the development of the classical Navier-Stokes equations, consisting of the
continuity, momentum and energy equations, are provided, however, taking most recent
research efforts into account, providing a summary of the knowledge that existed when
the author commenced his own derivations of the basic fluid flow equations in the
presence of density and temperature gradients in ideal gas flows. These derivations
resulted in a new set of equations, referred to as the extended Navier-Stokes equations.
The equations, thus derived in this thesis, were compared with other extended forms of
fluid flow equations based on completely different considerations for the causes of
differences between the classical and extended equations.

Subsequently, investigations were carried out for ideal gas flows through microchannels
and capillaries employing the extended Navier-Stokes equations, derived in this thesis.
The author’s initial work was on the numerical predictions of gas flows through
microchannels and excellent agreement was obtained with the experimental
measurements without invoking the Maxwell slip boundary condition at the solid walls.
Based on the insight obtained from the numerical simulations, he finally also helped a
group of researchers to derive a complete analytical solution to gas flows through
microchannels and capillaries. Based on this analytical solution procedure, it is not only
possible to obtain accurate description of the velocity distributions but also the pressure
distribution along the micro-conduit. More importantly, a characteristic pressure was
introduced based on the geometry and physical properties of the gas which was found to
describe the characteristics of the flow accurately. Further, accurate numerical
descriptions of microchannel flows with separation could also be provided based on the
vii
extended Navier-Stokes equations. A backward facing step flow was chosen as an
example to demonstrate the usefulness of the extended equations in predicting gas flows
with separation and excellent agreements were obtained with DSMC simulations.

Subsequently, numerical predictions of gas flows subjected to strong temperature


gradients were carried out and the prediction of one dimensional supersonic and
hypersonic shock waves in a monoatomic gas was chosen to demonstrate the usefulness
of the extended equations. Excellent agreement of computed shock structures with
experimental measurements was attained. Comparisons of various shock parameters
such as inverse density thickness, density asymmetry quotient and temperature – density
separation, were also provided and the extended Navier-Stokes equations were found to
perform satisfactorily. Further, the inadequacies of the classical and extended Navier-
Stokes equations in predicting hypersonic shock waves were stated. Interestingly, as
explained in this thesis, it was possible to provide physically meaningful explanations,
employing the extended equations, to some of the unsolved problems such as
thermophoresis and thermal transpiration. Further, it was evidently demonstrated that
one needed to consider the extended equations even in certain large scale gas flow
problems in order to obtain accurate and detailed description of flow and heat transfer
characteristics. Finally, the possibilities of further research work, employing these
powerful equations, were also identified.

viii
Table of Contents

ACKNOWLEDGEMENTS (v)

ABSTRACT (vii)

TABLE OF CONTENTS (ix)

1 INTRODUCTION AND AIM OF WORK 1

1.1 Background and overview 1


1.2 Outline of the thesis 7

2 THE NAVIER-STOKES EQUATIONS FOR FLUIDS WITH CONSTANT


PROPERTIES 11

3 THE EXTENDED NAVIER-STOKES EQUATIONS 18

3.1 Diffusion transport of mass 20


3.2 Diffusion transport of heat 22
3.3 Diffusion transport of momentum 19
3.4 The extended Navier-Stokes equations 24
3.4.1 Total velocity form of the extended equations 25
3.4.2 Extended total energy equation 26
3.5 Brenner’s extended Navier-Stokes equations 32
3.6 Comparison of the extended Navier-Stokes equations 34

4 NUMERICAL PREDICTIONS OF GAS FLOWS THROUGH


MICROCHANNELS 39

4.1 Introduction and literature review 39


4.2 Analytical solutions based on the CNSE with no-slip boundary condition 41
4.2.1 Derivations for microchannel flows 43
4.2.2 Derivations for capillary flows 46
4.2.3 Importance of diffusion mass transport 48
4.3 Governing equations, geometry and boundary conditions 50
4.4 Results and discussions 54
4.4.1 Velocity profiles 54
4.4.2 Pressure profiles 57
4.4.3 The ‘Knudsen-Paradox’ 58
ix
4.4.4 Comparison of mass flow rates 60
4.5 Insights into the physics of microchannel flows 61

5 ANALYTICAL TREATMENTS OF GAS FLOWS THROUGH MICRO-


CONDUITS 67

5.1 Introduction 67
5.2 Analytical solution procedure 68
5.2.1 Order of magnitude analysis 68
5.2.2 Analytical solutions of microchannel flows 72
5.2.3 Analytical solutions of capillary flows 74
5.2.4 Semi-analytical solutions of pressure in microchannels and capillaries
77
5.3 Results and discussions of gas flows through microchannels 78
5.3.1 Regimes of flows in microchannels 78
5.3.2 Discussions on pressure profiles 80
5.3.3 Discussions on pressure gradient profiles 84
5.3.4 Comparison of total mass flow rates 86
5.3.5 Discussions on the total velocity profiles 90
5.3.6 Conductance of the microchannel and the ‘Knudsen-paradox’ 93
5.4 Results and discussions of flows through capillaries 96

6. GAS FLOWS THROUGH COMPLEX MICROCHANNEL GEOMETRIES


101

6.1 Introduction 101


6.2 Comparisons of the ENSE solutions with Beskok [2001] and Celik and Edis
[2006] 104
6.2.1 Geometry and boundary conditions 104
6.2.2 Comparisons of results with Beskok [2001] 108
6.2.3 Comparisons of results with Celik and Edis [2006] 122
6.3 Comparisons of the ENSE solutions with Chakraborty [2010] 124
6.3.1 Geometry and boundary conditions 125
6.3.2 Results and Discussions for the case of Kno=0.01 126
6.3.3 Results and Discussions for the case of Kno=1 134

7. SOLUTIONS TO PROBLEMS WITH STRONG TEMPERATURE


GRADIENTS: SHOCK WAVES 143

7.1 Introduction 143


7.2 Computational inconsistencies faced in other methods 144
7.3 Geometry and boundary conditions 147
7.3.1 Variation of dynamic viscosity with temperature 150
7.4 Results and discussions 151
7.4.1 Density and temperature profiles 152
x
7.4.2 Velocity and Mach number profiles 158
7.4.3 Inverse density thickness, L1 163
7.4.4 Density asymmetry quotient, Q  165
7.4.5 Temperature-density separation,  T 167
7.4.6 Discussions on the definition of Knudsen number 168
7.4.7 Discussions on violation of laws of thermodynamics 170

8 IMPORTANT RESULTS AND SUGGESTIONS FOR FURTHER


RESEARCH 177

8.1 Introductory remarks 177


8.2 Outlook towards future research 180
8.2.1 Gas flows through microchannels 180
8.2.2 Computations of hypersonic shock structures 181
8.2.3 Thermophoresis 184
8.2.4 Thermal transpiration 187
8.2.5 Strongly heated pipe flows 191
8.3 Concluding Remarks

REFERENCES 195

APPENDIX 203

ZUSAMMENFASSUNG 215

xi
xii
Chapter 1

INTRODUCTION AND AIM OF WORK

1.1 Background and Overview


Fluid mechanics, one of the oldest and well-researched sciences, has wide spread
applications in many industries. The history of development of fluid mechanics and the
salient contributions from various eminent scientists is depicted in Figure 1.1. The basic
dynamic laws such as Newton’s laws of motion, Newton’s law of viscosity, Pascal’s
law etc. were developed in the 17th century. The development of these fundamental
laws had culminated in the formulation of the governing equations of fluid mechanics in
the nineteenth century, famously known as the Navier-Stokes equations, named after the
two distinguished scientists. These equations are referred to as the classical Navier-
Stokes equations (CNSE) in this thesis. The classical equations are believed to be
complete and capable of solving all kinds of fluid flow problems. This view point has
also been strengthened by decades of intense theoretical research contributions and their
validations with accurate measurements in sophisticated experiments. Hence, the
interests and focus of the modern fluid mechanical research community have primarily
been in the direction of developing more accurate measurement techniques and
implementing the various observed physical phenomena into computational fluid
dynamics software.

Figure 1.1 History of fluid mechanical research of various eminent scientists


[Durst (2008)]
Modern measurement techniques such as Laser Doppler Anemometry (LDA) are
capable of measuring even minute fluid velocities and turbulent fluctuations to a great
degree of accuracy. The ongoing development of high performance computing
machines and the availability of sophisticated computational fluid dynamics software,
armed with capabilities to handle complex geometries and novel grid generation
techniques, have revolutionized research and development activities in many fields of
engineering, science and medicine. Further, the development of turbulent models based
on the Reynolds averaged Navier-Stokes equations (RANS) has simplified the
treatments of complex turbulent flows of engineering significance without any
appreciable loss of accuracy. The development of modern transport systems such as
aircrafts and automobiles employs high-end sophisticated computational simulation
tools based on CNSE, as shown in Figure 1.2 and the agreement between the
computational results and experimental measurements has been exemplary. Further, the
computational and experimental measurement tools are widely employed as the
preferred choice in designing complex manufacturing systems with a great degree of
reliability and acceptability. Hence, one may say that this efficacy of the computational
fluid dynamics and measurement techniques in providing accurate answers to complex
fluid flow problems is one of the most significant achievements of the modern fluid
mechanics research.

Figure 1.2 Application of classical Navier-Stokes equations to engineering flow


problems [Durst (2008)]

On the contrary, there are certain problems in fluid mechanics that remain unsolved in
spite of the wealth of knowledge generated in these topics. The ‘slip-flow’ theory
2
employed in gas flows through micro-conduits and thermophoresis are just two of the
examples. The classical equations failed to predict either the characteristics of ideal gas
flows through microchannels at high Knudsen numbers as shown in Figure 1.3, see
Sambasivam and Durst [2011(b)], Arkilic et al [1997, 2001] and Gad-el-Hak [1997,
2002] or the physical reasons behind a phenomenon known as thermophoresis, a
process occurring very much within the continuum range, see Brenner [2005(a)-(d)].
The present state-of-the-art theories explaining the flow of gases through microchannels
and capillaries are replete with empiricism and lack sound theoretical basis. The
theoretical and numerical treatments of microchannels require a tuning parameter,
known as Tangential Momentum Accommodation Coefficient (TMAC), to match the
experimentally observed higher mass flow rates. The introduction of TMAC leads to
slip at the wall, i.e. the velocity of gas at the solid wall is not zero, and this is the only
possible way, as of today, one can match the experimentally observed mass flow rate of
an ideal gas through a microchannel with the theoretical or numerical analysis. The
bulk of the research in this field is devoted to tuning or fitting the results of theoretical
models based on the slip-velocity with the help of experimental measurements.

Experimental data
of Maurer et al
[2003]
Masss flow rate, Kg/s

CNSE

0.5Pin2  Po2 

Figure 1.3 Discrepancies of classical solutions and experimental measurements in gas


flows through microchannels

However, it can be shown categorically that the typical roughness values of


microchannel walls will result in the no-slip boundary condition at the wall; see Mo and
Rosenberger [1990]. Furthermore, prediction of flow characteristics based on the ‘slip-
flow’ theory becomes difficult, rather impossible, as soon as a simple complexity in
geometry, say a backward facing step, is introduced. The prevailing conclusion of the
research community is that CNSE do not give accurate predictions for flows through
microchannels in the slip-flow regime and beyond, i.e. for Knudsen number values
more than 0.01. Further, at Knudsen numbers beyond 0.1, it is generally considered that
3
CNSE even with the slip boundary condition cease to be valid; see Zhang [2005].
Therefore, the Direct Simulation Monte Carlo (DSMC) or Lattice Boltzman Modelling
(LBM) techniques are employed to numerically predict the flow of ideal gases through
microchannels and capillaries at larger Knudsen numbers. In short, it may be said that
the sound physical understanding of flow of gases through complex network of
microchannels, a basic requirement for the ongoing development of Micro-Electro-
Mechanical Systems (MEMS), still remains evasive.

CNSE

Figure 1.4 Discrepancy in computed shock structure employing classical equations and
experimental measurements by Alsmeyer [1976]; Inlet Mach number 3.81

The other inadequately treated phenomenon, known as thermophoresis, is defined as the


movement of small non-Brownian particles suspended in a gas from hotter to colder
regions without any perceptible motion of the fluid. Interestingly, this phenomenon
occurs well within the continuum regime, i.e. the fluid can be accurately defined and
treated as continuum instead of molecules or atoms, and yet CNSE cannot provide any
valid explanation to this phenomenon; see Brenner [2005(a)-(d)]. Once again, the
explanations provided by other prevailing theories are based on empirical parameters to
match experimental observations and questionable thermal slip assumptions at the
surface of even a microscopic particle.

To further substantiate the incompleteness of the basic equations of fluid mechanics, the
observed discrepancy between experimental measurements and numerical predictions of
the structure of shock waves is depicted in Figure 1.4; see Alsmeyer [1976] and
Greenshields and Reese [2007(a) and (b)]. According to Greenshields and Reese
[2007(a)], the numerically computed spread of the shock wave employing CNSE is less
than the experimental measurements. In order to get good agreement with the
experimental measurements, it is required to employ extensions to the classical
equations with higher order terms, for example Burnett equations. It is essential to note
4
that the extended hydrodynamic models, available as of today, are based on
mathematical perturbations rather than derivations based on physics based arguments
and hence suffer from convergence difficulties and oscillations. More importantly, it is
well known that the shock structures obtained with such extended hydrodynamic
models violate the second law of thermodynamics.

40
CNSE
35 Experiment
Local Nusselt number

30

25

20

15

10

5
0 5 10 15 20 25
Nondimensional axial distance

Figure 1.5 Discrepancy in computed local Nusselt number employing classical


equations and experimental measurements by Shehata and McEligot [1998]

The classical equations fail not only in the fringe regions of the near-continuum or
moderately-high Knudsen number flows but also in certain large scale engineering
systems of very small Knudsen numbers. For example, the local Nusselt number values
of a strongly heated turbulent pipe flow obtained by CNSE are compared with the
experimental measurements of Shehata and McEligot [1998] in Figure 1.5. It can be
observed that the classical RANS equations with k    turbulence model failed to
predict the heat transfer characteristics with a large discrepancy with the experimental
measurements. In general, the inadequate modelling of buoyancy driven turbulence
production and dissipation terms will be suspected to be the reason for the failure of
RANS based turbulence model in predicting the flow characteristics properly. Hence,
based on the experimental results, the turbulence model constants will be adjusted so
that good agreements between the numerical predictions and experimental
measurements are achieved. However, it can also be argued that such large
discrepancies may not happen because of poor modelling of buoyancy driven
production of turbulent kinetic energy alone and it can also be because of some other
physical phenomenon which has not been incorporated in the classical governing
equations, as shown later in this thesis.
5
From the above mentioned examples, discussed based on Figures 1.3–1.5, one may
conclude that CNSE are not capable of handling every ideal gas flow problem
accurately or in other words, the classical equations are, in some respect, incomplete.
Obviously, one is expected to get disturbed by the aforesaid two contrasting view points
of the present-day status of fluid mechanics. On one hand, we have a thorough well-
developed system of equations, aptly supported by the developments of experimental
and computational tools that help to solve complex flow situations with a great degree
of accuracy. On the other hand, the classical equations struggle to provide appropriate
answers to some of the intriguing problems of practical interest that remain unsolved for
decades. The aim of this thesis is to develop physically sound answers to some of the
above mentioned problems by revisiting the derivations of CNSE.

In CNSE, it is customary to introduce the constitutive relationships proposed for the


molecular momentum and heat transfer given by the following equations.
 U j  2
Molecular momentum transfer  ij     U i     ij U k (1.1)
 x i  3
 x j x k
T
Molecular transport of heat q i   (1.2)
xi

Equations (1.1) and (1.2) were derived from the Stoke’s hypothesis of molecular
momentum transport based on the Newton’s law of viscosity and Fourier’s law of heat
diffusion, respectively. These relationships are considered to be applicable to all
Newtonian fluids and the coefficients in the constitutive relationships, dynamic
viscosity  and thermal conductivity  are modelled as functions of temperature.
However, Durst et al [2006] proposed an extension to CNSE during the period when the
research work for this thesis was commenced. The derivations of the extended Navier-
Stokes equations (ENSE) were revisited as part of the research work summarized in this
thesis and finally, the extended equations were obtained in the present form discussed in
this thesis. Moreover, the total velocity, defined as the sum of convective and diffusion
velocities, based formulation of the Navier-Stokes equations was also derived. This
form of the equations was found to be easy to employ in obtaining physically
meaningful solutions to gas flows through closed conduits. Further, Brenner [2005(a)-
(c)] also proposed another set of extensions to the Navier-Stokes equations to explain
the thermophoretic motion by altering the classical constitutive relationships for the
molecular momentum and heat transfer. On comparing with the extended equations
suggested by Brenner, it is stressed here that the extended equations derived in this
thesis was found to provide physically-correct explanations for the flow of ideal gases
in the presence of temperature and density gradients. In this thesis, the solutions
obtained with ENSE for a number of problems are described in detail and the results
were found to have excellent agreement with the experimental measurements.

6
1.2 Outline of the thesis

The thesis is structured in the following way. In Chapter 2, the well-known derivations
of CNSE are presented. Though the derivations of the classical constitutive
relationships, given by equations (1.1) and (1.2), are available in many standard text
books on the subject, it was felt that the derivations of the classical equations could be
used as a starting point to obtain the extended form of the equations discussed later in
the thesis. In Chapter 3, it is further argued that the discrepancies observed between the
experimental and theoretical/numerical predictions of certain ideal gas flow problems
directly point towards the incompleteness of CNSE. It is shown that CNSE are valid
only for flow fields free of mass diffusion, i.e. gas flows without density and
temperature gradients. Thereafter, it is evidently established that when density and
temperature gradients are present in ideal gas flows, CNSE need to be extended by
incorporating self-diffusion transport of mass and the allied additional diffusion
transport of heat and momentum. Furthermore, the concept of ‘total velocity’, defined
as the vector sum of convective and diffusion velocities, is, thereby, introduced and the
derivations of the extended equations based on the total velocity formulation are also
described in chapter 3. Furthermore, the extended Navier-Stokes equations proposed by
Brenner [2005(a)] are also presented and the observed discrepancies between the two
sets of equations are elaborated. It is argued that ENSE derived in this thesis are
complete and have a strong physical basis and hence, it can be employed in solving gas
flow problems with density and/or temperature gradients.

As the first application of ENSE, numerical solutions of ideal gas flows through
microchannels are presented in chapter 4. The available literature on gas flows through
microchannels clearly concluded that CNSE failed to determine the characteristics of
gas flows through micro-conduits in the alleged ‘slip-flow’ regime since the
experimentally measured mass flow rates through the conduits were found to be more
than those predicted by the theoretical/numerical analysis. As it could be observed from
the published literature, the ‘slip-flow’ theory was introduced in the theoretical analysis
of microchannel flows based on only the above mentioned observation and the
determination of the ‘slip coefficient’ employing experimental measurements based on
the ‘Maxwell-slip’ theory was mandatory for any plausible theoretical analysis.
However, it can evidently be shown that the roughness of solid surfaces employed in
manufacturing micro-conduits will necessitate the no-slip boundary condition to be
satisfied and hence the introduction of the Maxwell slip velocity, applicable only in the
case of molecularly smooth walls, in the theoretical/numerical analysis is empirical and
lacks sound theoretical foundation. Further, the introduction of minor complexity in the
geometry, such as a backward facing step, will cease the usability of the slip flow
theory.

Before employing the ENSE to solve gas flows through micro-conduits, it was required
to estimate the importance of the acceleration terms in the momentum equations. In the
published literature, there are analytical solutions for fully-developed gas flows through
microchannels with both the no-slip and the Maxwell-slip boundary conditions. In this
7
chapter, analytical solutions based on integral analysis of CNSE including the non-
linear acceleration terms with the no-slip boundary condition for microchannels and
capillaries are obtained. It has been found that the effect of the non-linear inertial terms
can be neglected and gas flows through micro-conduits can be treated almost like fully-
developed flows. Further, the numerically predicted results obtained with the ENSE
have been found to agree excellently with the experimental measurements of gas flows
through microchannels. It is claimed that this conclusively proves the physical
mechanism causing the additional mass flow rate through microchannels to be the self-
diffusion of mass driven by density gradients. It can also be observed in this chapter
that the numerical predictions employing the ENSE are able to determine each and
every characteristic of gas flows through microchannels without making any empirical
or modelling assumptions, even at high Knudsen numbers of O 1 . Further, phenomena
such as the ‘Knudson paradox’, i.e. the presence of a minimum in the conductance
curve, are also predicted and explained with ease.

It was felt that the insights obtained from numerical simulations of gas flows through
microchannels could lead to developing an analytical solution, as shown in chapter 5.
The extended equations in total velocity form were employed to develop the analytical
solutions for microchannels and capillary flows and the results agree very well with the
experimental measurements. Further, the introduction of characteristic pressure has
been observed to be an invaluable tool in the analysis of gas flows through micro-
conduits. Furthermore, a semi-analytical solution procedure was also developed to
obtain the pressure distribution along microchannels and capillaries and the results also
agree well with the experimental measurements and numerical predictions alike.
Further, the presence of complex geometries in micro-conduits limits the usefulness of
the slip-flow theory. Since the slip-velocity at the solid wall needs to be determined
from the experimental measurements, the complex geometry, say a sudden change in
the cross-section, limits the application of the slip-flow theory. On the contrary,
treatment of complex geometry did not pose any difficulty in the analysis with the
ENSE, as shown in chapter 6 for the case of prediction of flow characteristics of
microchannels with a backward facing step. Based on the results presented here, it is
possible to conclude that the problem of gas flows through micro-conduits of any
arbitrary shape can be solved by employing the ENSE.

As stated in the previous subsection, the classical equations fail to predict the
hypersonic shock structures accurately and the presence of non-continuum effects at the
supposedly high Knudsen number is stated to be the reason for the failure of the CNSE
in predicting such strongly compressible flows. In chapter 7, the structures of one
dimensional supersonic and hypersonic shock waves predicted employing the ENSE
have been compared with experimental measurements and the profiles agree
exceedingly well. The modified extended equations proposed by Brenner [2005(a)]
were found to fail at higher inlet Mach number values. Further, shock parameters, such
as inverse density thickness and density asymmetry quotient, predicted by the ENSE
also agree very well with experimental measurements and the temperature-density
separation distance has been found to be under-predicted by the ENSE in comparison
8
with the DSMC simulations. Moreover, in this chapter, some important discussions on
the choice of characteristic length in defining Knudsen number in shock waves and
observed thermodynamic violations in shock wave simulations are also presented.

In chapter 8, the summary and conclusions of results obtained for all the problems
considered in this thesis are provided. Further, the possibilities of employing the ENSE
in solving a number of flow cases of physical significance are also outlined along with
sample results. These include the descriptions of thermophoretic motion and thermal
transpiration and heat transfer from a strongly heated pipe to a gas. Further, the
agreement of numerical predictions of shock waves employing the ENSE with
experimental measurements has been found to improve when the self-diffusion
coefficient is modified based on the Chapman-Enskog theory. It is finally stated that
the ENSE has been found to be a valuable tool in providing physically sound answers to
a number of unsolved gas flow problems.

9
10
Chapter 2

THE NAVIER-STOKES EQUATIONS FOR


FLUIDS OF CONSTANT PROPERTIES
As shown in a number of standard text books on fluid mechanics, see Durst [2008],
Bird Stewart and Lightfoot [1960] and White [1974], the classical governing equations
of fluid mechanics for an isothermal ideal gas flow can be written in the following
way.

Continuity equation:
  U iC

 0
 (2.1)
t xi
 U Cj   U iCU Cj  C
P  ij
Momentum equation:     g j (2.2)
t xi x j xi
Equation of state:   P T (2.3)

In the above mentioned equations, U iC is the convection velocity,  the local density,
P the local static pressure, T the static temperature,  the gas constant,  ijC the classical
molecular momentum transport terms and g j the acceleration due to gravity in the j
direction.

Assuming isothermal flow conditions and taking into consideration the symmetry of
the term  ijC , i.e.  ijC   Cji , it can be observed that there are eleven unknowns in the
above questions, namelyU 1C , U 2C , U 3C , P ,  , 11C , 12C , 13C , 22C , 23
C C
and  33 . For these
unknowns, there are only five partial differential equations, in equations (2.1) to (2.3),
available. Hence, this is an incomplete system of equations and it is therefore
necessary to state additional equations, i.e. to express the unknown terms  ijC in a
physically well-founded manner, as functions of the velocity components. For the
whole class of fluids whose molecular momentum transport properties can be
classified as ‘Newtonian’, the constitutive relationships can be obtained from the
kinetic theory of gases, see Durst [2008]. In the standard text books, the velocity
components and molecular momentum transfer tensor do not have any superscripts
and are denoted by U i ,U j and  ij , respectively. However, the superscripts are used in
this thesis to differentiate the convective velocity from the total and diffusion
velocities introduced in the following chapter.

Considering a control volume in the space as shown in Figure 2.1, with all boundaries
parallel to the Cartesian coordinate system, the j-momentum transported in the i-
direction by an instantaneous velocity field Û iC can be stated as follows:
Iij   Uˆ iC Uˆ Cj Fi (2.4)
The instantaneous velocity components Û iC and Û Cj can be decomposed into their
respective velocity components of the flow U iC and U Cj and the molecular velocities
u iM and u Mj , respectively and hence, equation (2.4) can be rewritten as:

Figure 2.1 j-Momentum input due to flow through the plane [Durst (2008)]

Uˆ iCUˆ Cj   U iC  u iM U Cj  u Mj    U iCU Cj  u iM U Cj  u Mj U iC  u iM u Mj  (2.5)

By time averaging, the time-averaged total momentum change of the volume can be
obtained as:
 
Uˆ iCUˆ Cj   U iCU Cj  uiM U Cj  u Mj U iC  uiM u Mj  (2.6)
     
 Term - I Term - II Term - III Term - IV 

The total momentum input, given by equation (2.6), consists of four terms as given
below:
Term I: -Momentum input in the direction due to the velocity field U iC of the fluid.
Term II: j-Momentum input in the direction due to the time-averaged molecular
velocity in the direction u iM .
Term III: -Momentum input in the direction due to the time-averaged molecular
velocity in the direction u jM .
Term IV: For i  j , u iM u Mj  0 as the molecular motion in the three coordinate
directions are not correlated. Further, when i=j, the term u iM u Mj results in
the time-averaged molecular motion caused pressure.

12
In equation (2.6), Term I represents the convective transport of j-momentum in the i-
direction due to the convective velocityU iC . In gases, the molecular mean free path l is
of finite dimensions, i.e. ≠ 0, and for this reason the time averages u iM u Mj and u Mj u iM
are not equal to zero when i=j. The diffusive transport of momentum in the i and j
directions are given by Terms II and III in equation (2.6). In order to calculate these
two terms, the following considerations have been carried out. When x1  x2  x3  a
the number of molecules moving in the direction and passing the plane in Figure
2.2 in the time duration t , can be expressed as:
1
zi  na 2uiM t (2.7)
6
where n is equal to the number of molecules per unit volume, a 2 the magnitude of the
area , and u iM the mean velocity of the molecules in the direction. Further, one
can also write the following expression for the mass transport through as follows:

l
l

Figure 2.2 j-Momentum input in the xi direction caused by the mean molecular velocity
u iM [Durst (2008)]

1
mM zi  
mM n a 2 u iM t (2.8)
6 

where m M represents the mass of a molecule and thus the density can be expressed as
  m M n . Here, it is necessary to note that the mean molecular velocity u M is constant
in all the directions, i.e. u M  u iM  u jM . However, the subscripts i and j are used for the
mean molecular velocity to denote only the motion in that particular direction.

As shown in Figure 2.2, let us consider two planes located at a distance ± above and
below the plane Fi . Further, the velocity components in the j direction in these two
planes are given by U Cj ( xi  l ) and U Cj ( xi  l ) . The -momentum inputs in the positive
and negative directions arriving at Fi can be stated as:
Iij   z i m M U Cj ( xi  l )
(2.9)
13
Iij   z i m M U Cj ( xi  l ) (2.10)

Therefore, the net momentum input can be expressed as the vector sum of the values
given in equations (2.9) and (2.10) as given below:
Iij  z i mM U Cj xi  l   U Cj  xi  l  (2.11)

By incorporating the expression given for z i in equation (2.7), equation (2.11) can be
rewritten as:
1
Iij   
mM n a 2u M t U Cj xi  l   U Cj xi  l 
6 
 (2.12)

The net momentum input per unit area and unit time can be obtained by Taylor series
expansion of the velocity terms around x i . Neglecting higher order terms, equation
(2.12) can be expressed as:
1 Iij 1 M  C U Cj U Cj 
Term II   ijI   u U 
 j i x   l  U C
j  x i   l (2.13)
a 2 t 6  xi xi 
Further, when one substitutes the expression of dynamic viscosity given by the kinetic
theory of gases, equation (2.13) can be expressed as follows:
C C
1 U j U j
 ijI   u M l   (2.14)
3  xi
 xi

Analogous to this, it is possible to carry out considerations to derive the expression for
Term III in equation (2.6). The number of molecules moving in the j-direction can be
written as:
1
z j  na 2u jM t (2.15)
6
Subsequently, the j-momentum inputs from the planes located at distances  l from the
plane under consideration B shown in Figure 2.3, can be expressed as given below:
I ji   z j m M U iC ( x j  l )
(2.16)
I ji   z j m M U iC ( x j  l ) (2.17)
Therefore, the net momentum input can be expressed as:

Iij  z j m M U iC x j  l   U iC x j  l  (2.18)

Analogous to the derivations shown in equations (2.12) to (2.14), the following


equation can be obtained by simplifying equation (2.18).
II 1 M U iC U iC
Term III   ij   u l   (2.19)
3  x j
 x j

14
B   Fj

B   Fj

l l

Figure 2.3 i-Momentum input in the x j direction caused by with molecular velocity
u jM [Durst (2008)]

Hence, the total molecular transport of momentum in the control volume can be
written as given below:
C I II
 U Cj U iC 
   
ij ij ij      (2.20)
 xi x j 

Equation (2.20) provides the total momentum input  ijC when  =constant, i.e. when
d V   0 . However, when   constant, an additional term needs to be added to
dt
C
 ij shown in equation (2.20) which is caused by the volume increase of the control
volume. For the volume increase of a control volume at point xi and time , see Durst
[2008], it can be written as:
d V  U iC
 V  (2.21)
dt xi

Further, the corresponding surface increase can be given as:


d F  2 U iC
 F  (2.22)
dt 3 xi

With this increased surface area, an increased momentum input results which can be
expressed as given below:
2 U kC
 ijIII     ij (2.23)
3 x k

This term has to be added to equation (2.20) in order to obtain the general  ijC
relationship for the total momentum input per unit time and unit area for ideal gases as
given below:

15
 U Cj U iC  2 C
  
C
ij     ij  U k (2.24)
 xi x j  3 xk
 

When equation (2.24) is considered for ijC , the following closed system of basic
equations of fluid mechanics can be obtained for isothermal flow conditions and
Newtonian fluids.
  U iC 
Continuity equation:  0 (2.1)
t xi
 U j   U iCU Cj 
C C
P  ij
Momentum equation:     g j (2.2)
t x i x j x i
Equation of state:   P T (2.3)
C
 U Cj U iC  2 U kC
Molecular momentum transport: ij        ij (2.24)
 xi x j  3 xk

When non-isothermal conditions are encountered, it is necessary to incorporate the


thermal energy equation (2.25), into the system of equations shown in equations (2.1)
- (2.3) and (2.24) and the thermal energy equation is given below:
 e   U iC e  q i U Cj C
U Cj
  P   ij (2.25)
t x i xi x j x i
In equation (2.25), the internal energy e is given by e  CV T where CV is the specific
heat at constant volume. The unknown term q i for the molecular transport of heat is
closed by the Fourier’s law of conduction as:
T
q i  q iF   (2.26)
xi
where  is the local thermal conductivity of the fluid.

Equations (2.1)-(2.3) and (2.24)-(2.26) have successfully been employed in solving a


number of ideal gas flow problems in the past. However, it is stressed that when strong
density and temperature gradients are encountered in the flow, an important transport
mechanism, i.e. self-diffusion of mass, is expected to assume significance and this is
completely neglected in the classical governing equations mentioned above. When the
extent of self-diffusion of mass is negligible, the classical equations do not pose any
serious errors and have been used successfully to predict non-isothermal gas flow
problems in the past. However, there are gas flow problems with strong density and
temperature gradients of practical interest where the self-diffusion of mass becomes
important. Therefore, the self-diffusion transport of mass was incorporated into the
classical Navier-Stokes equations and the extended equations were obtained to solve
gas flow problems with strong temperature and density gradients, as discussed in the
subsequent chapters.

16
Chapter 3

THE EXTENDED NAVIER-STOKES


EQUATIONS
When the classical Navier-Stokes equations (2.25) to (2.30), also referred to as the
CNSE, were employed to solve ideal gas flow problems such as flow through micro-
conduits and prediction of shock waves, one could observe discrepancies between the
experimental measurements and theoretical and/or numerical predictions, as shown in
Figure 1.1. In order to address such discrepancies, extensions to the CNSE have been
suggested, for example Durst et al [2006] and Brenner [2005(a)]. Extensions to the
Navier-Stokes equations are needed when flows of ideal gases are solved in the
presence of strong density and temperature gradients. Brenner [2005(a)] proposed a set
of extended equations in order to describe the thermophoretic motion in gases, and these
are provided at the end of this chapter for comparisons with the extended equations
derived in this thesis.

Durst et al [2006] argued that when strong density and/or temperature gradients are
present in the flow, the influence of self-diffusion of mass cannot be neglected, as
explained in detail in subsequent sections. The presence of density and temperature
gradients in ideal gas flows results in an additional molecular diffusive transport of
mass caused by the self-diffusion in the direction opposite to the gradients of density
and temperature. Further, it was also claimed that the said mass diffusion also gave raise
to additional heat and momentum diffusion. Hence Durst et al [2006] suggested that the
classical constitutive relationships for the diffusive transport of heat and momentum,
given in equations (2.24) and (2.26), also had to be enhanced with the additional terms
caused by the self-diffusion of mass. When the effects of self-diffusion of mass are
incorporated in the CNSE, one can obtain the extended Navier-Stokes equations as
described in this chapter, also referred to as the ENSE. The equations derived in this
thesis are slightly different from those presented by Durst et al [2006], but the basic
derivation procedure remains the same.

In this chapter, expressions for the extended diffusion transport of mass, momentum and
energy are derived. Further, the extended constitutive relationships obtained for the
diffusion transport are incorporated in the CNSE to obtain the ENSE. As a special case,
the extended equations are also obtained in the total velocity form and this form was
found to be useful in certain flow simulations with solid boundaries. The extended
equations suggested by Brenner [2005(a)], referred to as the BNSE, are also provided
for comparison.
3.1 Diffusion transport of mass

Usually, the diffusion transport of mass in gases is considered in the classical


governing equations, shown in equations (2.1)-(2.3), when multiple species are present
in the problem under consideration. This well-known problem is usually referred to as
multi-component diffusion. However, diffusion transport of mass also occurs in gases
when only one gas is present, and this is known as self-diffusion. Self-diffusion is
always neglected in gases with the presumption that its influence is insignificant in
comparison with the convective transport of mass. However, in the presence of strong
density and temperature gradients, one can easily demonstrate that the self-diffusion of
mass cannot be neglected and at times can attain magnitudes of the order of
convective mass transport. Therefore, it is argued that incorporation of the self-
diffusion transport of mass in the classical continuity equation (2.25) is mandatory for
consistent treatment of ideal gas flows.

T xi  l 
  x  l 
 i
 M
u xi  l 
U C x  l 
 j i

T xi  l 
l  
l  xi  l 
 M
u xi  l 
U C x  l 
 j i

Figure 3.1: Self-diffusion mass transport due to density and temperature gradients
[Durst (2008)]

To analyze the molecular transport processes, one can consider Figure 3.1. The self-
diffusion transport of mass can be estimated from two planes separated by a distance l
from the plane under consideration. Applying the kinetic theory of gases, the
1
molecular transport of mass can be written as m xDi l   xi  l  u M xi  l  and
6
1
m xDi l    xi  l  u M xi  l  from the planes located at  xi  l  and  xi  l  , respectively.
6
Here  is the local density of the gas and u M is the molecular mean velocity, defined
as:
18
8T
uM  (3.1)
mM
where T is the absolute temperature,  the Boltzmann constant, mM the molecular
mass and l the local molecular mean free path of the considered ideal gas.

The net diffusive mass flux can be expressed as:


1

m iD   xi  l u M  xi  l     xi  l u M xi  l 
6
 (3.2)
One can clearly observe from equations (3.1) and (3.2) that when there are no spatial
density and temperature gradients in the flow field, the net diffusive mass flux m iD  0 .
However, when density and temperature gradients are present, the net diffusive mass
flux m iD is present and it can be evaluated based on the derivations given below.

When equation (3.2) is expanded using Taylor series, one can obtain the following
expression after truncating the higher order terms:
M M
1       
m iD     xi    l  u M xi   u  l     xi    l  u M  xi   u l   (3.3)
6  x i   xi   x i   x i 
Taking into consideration the isotropy of the molecular motion and by including only
the product terms that contain first-order derivatives, the diffusive mass transport in
the xi direction can be written from equation (3.3) as:
1   u M 
m iD   l u M   (3.4)
3  xi xi 

The diffusion coefficient D can be written as:


D  1 lu M
3
  (3.5)

By substituting equation (3.5), equation (3.4) can be rewritten as:


D    u M 
m i   D   M  (3.6)
 xi u xi 

Further, substituting   D for ideal gases and employing equation (3.1), the
following form can be obtained:
   T   1  1 T 
m iD   D      
D    (3.7)
 x i 2T x i   
 x i 2T x i 

Equation (3.7) shows that the diffusive transport of mass is driven by both the local
density and temperature gradients. Further, assuming local thermodynamic
equilibrium, from the ideal gas law given by equation (2.3), one can write:
1  1 P 1 T
P   T    (3.8)
 xi P xi T xi

19
Substituting equation (3.8) in equation (3.7), one can also obtain the following:
 1  1 T   1 P 1 T 
m iD   
D         (3.9)
   x i 2T x i   P x i 2T x i 

The results of the above derivations are well known and they yield the following
terms:

Fick’s mass diffusion term: m iF   D (3.10)
xi
 D  T
Soret’s mass diffusion term: m iS    (3.11)
 2T  xi
where D S   2T  is the Soret diffusion coefficient.

Hence it can be seen that the derivations proposed in this section have only yielded
well-known relationships for the diffusive mass fluxes in gas flows and not something
unknown to the fluid mechanical research community. However, the terms in equation
(3.9) are not taken into account in the conventional theoretical or numerical analysis of
ideal gas flows, since they are considered negligible with respect to the convective
mass transport terms in the continuity equation (2.1).

3.2 Diffusion transport of heat

One can clearly observe from Figure 3.1 that for every component of the diffusive
transport of mass, a corresponding diffusive transport of heat is associated. Hence one
can write the following equations using quantities for the mean molecular motion:
1 1
q xDi l   xi  l u M xi  l exi  l  and q xDi l    xi  l u M  xi  l e xi  l  (3.12)
6 6
where e is the local internal energy of the gas and q xDi l and q xDi l are the molecular heat
transport from the xi  l  and  xi  l  planes, respectively. Further, for an ideal gas,
the internal energy e can be written as e  CV T , where CV is the specific heat at
constant volume. Now, equation (3.12) can be rewritten as:
C C
q xDi l  V  xi  l u M xi  l T xi  l  and q xDi l  V   xi  l u M  xi  l T  xi  l  (3.13)
6 6

Similar to equation (3.2), one can write for the net diffusive heat flux:
C
 
qiD  V   xi  l u M xi  l T  xi  l     xi  l u M xi  l T  xi  l 
6
(3.14)

Further, by a Taylor series expansion of the terms in square brackets in equation (3.14)
and truncating the series after the first-order derivatives, the following expression can
be obtained:
20
   M u M  T  
   x i    l  

 u  x i    l  
 
 T  x i     l  
  
D CV  xi  xi  xi   (3.15)
q i  
6    M M
 
   xi   l  u  xi   u l  T  xi   T l 
  xi  xi  xi 

Equation (3.15) can be rewritten, neglecting all terms that contain products of two
first-order derivatives or more, as:
C  T  T u M T 
q iD   V lu M    M   (3.16)
3   x i u x i x i
M
From the definition of u M given by equation (3.1), one can derive u x as:
i

u M 8 1 T
 (3.17)
xi mM 2 T xi
and subsequently the following relationship can be obtained:
1 u M 1 T
M
 (3.18)
u xi 2T xi

Substituting equation (3.18) in equation (3.16), the following expression for qiD is
obtained:
C T  1  1 T 1 T 
q iD   V lu M     
3   xi 2T xi T xi 
  (3.19)
 1 M   1 T 1  1 T 
 CV T  l u      
3   T xi  xi 2T xi 
 D 
Equation (3.19) can be rewritten as:
  T  1  1 T 
q iD   CV D   CV T D   (3.20)
   xi   xi 2T xi 
     
 m iD

Hence one can obtain the final equation for the total diffusive heat transport in an ideal
gas as:
T
q iT     m iD CV T (3.21)
xi   
 qiSD
qiF

which consists of the following.

 T 
Fourier’s heat diffusion term: q iF     [3.22, same as equation (2.26)]
 xi 

21
An additional heat diffusion term as a result of the self-diffusion of mass:
q iSD  m iD C V T (3.23)

Interestingly, one can also interpret equation (3.20) differently as follows:


3 T    T 
q iD     CV T D    eff  Df (3.24)
2 xi xi 
  xi xi

   
Temperatur e-driven Density -driven heat
heat transport term transport term  Dufour - term

The Dufour term, in equation (3.24) is often considered in treatments of binary


diffusion problems, see Coelho and Silva Telles [2002], but not for “self-diffusion”
problems. Furthermore, one can observe in equation (3.24) the change in the effective
3
thermal conductivity eff   , whereas the Dufour diffusion coefficient D f depicts
2
its dependence on the temperature, T.

3.3 Diffusion transport of momentum

From Figure 3.1, one can write the following expressions for the diffusive transport of
momentum from the two planes:
1 1
I xDi l   xi  l u M xi  l U Cj xi  l and I xDi l    xi  l u M  xi  l U Cj  xi  l  (3.25)
6 6
where U Cj is the local convective velocity in the j direction and I xDi  l and I xDi  l are the
molecular momentum transport from the xi  l  and  xi  l  planes, respectively.

Further, one can write the net diffusive momentum flux of momentum j in the i
direction as:
1

 ij   xi  l u M xi  l U Cj xi  l     xi  l u M  xi  l U Cj xi  l 
6

(3.26)

Further, by expanding the terms in square brackets in equation (3.26) employing


Taylor series and truncating the series after the first-order derivatives, the following
expression can be obtained:
   M u M  C U Cj  
   xi    l  u xi    l U j  xi    l   
1   x i    x i   x i  
 ij    (3.27)
6    M u M  C U Cj 
    xi   l  u xi   l U j xi   l 
  x i  x i  x i 

Neglecting all terms that contain products of two or more first-order derivatives,
equation (3.27) can be rewritten as:

22
C C C
1 M U j  U j u M U j 

 ij   lu   
 M   (3.28)
3   xi u xi xi 

By substituting equation (3.18), equation (3.28) can be rewritten as:


 
1   1  1 T
C
1 U j 
C M
 ij  U j  lu      C  (3.29)
3
      x i 2T x i U j x i 
  

Further, equation (3.29) can be rearranged as:


U Cj  1  1 T 
 ij     U Cj    (3.30)
x i   x i 2T x i 

Employing equation (3.9), the final form of equation (3.30) can be expressed as:
U Cj
 ij     m iDU Cj (3.31)
xi

Similarly, one can also write the expression for the net diffusive flux of momentum i
in the j direction as:
1
 
 ji   x j  l u M x j  l U iC x j  l    x j  l u M x j  l U iC x j  l 
6
(3.32)

Following the steps mentioned earlier, one can obtain the following expression:
U iC
 ji     m DjU iC (3.33)
x j

Since the flow is considered to be compressible, the expansion in time of the fluid
element yields an increase in momentum input to the fluid element. Therefore, similar
to the derivations in equations (2.20) and (2.21), one can write the following
expression:
2  1  1 
 ijVD   ij   x k  l u M  xk  l U kC  x k  l     x k  l u M x k  l U kC  xk  l  (3.34)
3  6  6 

Expanding the terms in parentheses into a Taylor series and manipulating equation
(3.34) as in previous steps, one can obtain the following momentum transport tensor:
2 U kC 2  1  1 T 
 ijVD   ij    ij U kC    (3.35)
3 xk 3   xk 2T xk 

Employing equation (3.9), equation (3.35) can be rewritten as:


2 U kC 2
 ijVD   ij    ij m kDU kC (3.36)
3 x k 3
23
Combining equations (3.31), (3.33) and (3.36), one can obtain the complete
constitutive diffusive momentum transport tensor as:
 U Cj U iC  2 C
 T
ij        ij  U k  m iDU Cj  m DjU iC  2  ij m kDU kC (3.37)
 xi x j  3 xk 3
 

This expression for the total momentum diffusion due to molecular action  ijT contains
the classical momentum diffusion terms given in equation (2.24) and also three
additional terms arising due to the self-diffusion transport of mass. It is interesting
that the well-known derivation procedure presented in this section yields additional
terms when one takes property variations into account.

3.4 The extended Navier-Stokes equations

When one considers the extended diffusion transport terms shown in equations (3.9),
(3.21) and (3.37) as the constitutive relationships in the CNSE presented in equations
(2.1), (2.2) and (2.25), the ENSE are obtained as given below. When the self-diffusive
mass transport, equation (3.9), is included in the classical continuity equation, shown
in equation (2.1), one can obtain the extended continuity equation:
 U iC  m D
  i (3.38)
t xi xi
Equation (3.38) shows that only in those flows where the molecular mass diffusion m iD
is negligible in comparison with the convective mass transport can the classical
continuity equation be obtained.

Similarly, when the extended constitutive relationship for the molecular momentum
transport, equation (3.37), is used in the classical momentum equations, equation (2.2),
one can obtain the extended momentum equation as:
  U Cj U C  2 U kC 
    i 
  ij  

 U Cj   U C
i U Cj  
P

   xi x j  3 x k   g (3.39)
j
t xi x j xi  2

  m i U j  m j U i   ij m k U k 
D C D C D C
 3 
Equation (3.39) can also be rewritten as:

 U Cj    U C
i U Cj    P   T
ij
 g j (3.40)
t x i x j x i

Similarly, substituting the extended molecular energy transport term, equation (3.21),
and the extended molecular momentum transport terms, equation (3.37), in the
classical energy equation (2.25), one can obtain the extended energy equation as:
 e   U iC e  q iT U iC T
U Cj
  P   ij (3.41)
t xi xi xi xi
24
It is stressed that the set of extended Navier-Stokes equations, given by equations
(3.38), (3.40) and (3.41), must be solved to obtain accurate and physically meaningful
solutions when strong temperature and density gradients are present in gas flows.
Further, in some special flow geometries with solid walls, such as straight
microchannels, it may be helpful to express the extended Navier-Stokes equations in
the total velocity form, as shown in the following section.

3.4.1 Total velocity form of the extended equations

One can express the self-diffusive mass transport m iD as a function of diffusion velocity
which is defined as:
m iD   U iD (3.42)
where U iD is the diffusion velocity in the i-direction. Employing equation (3.9), one
can express the diffusion velocity U iD as:
  1  1 T    1 P 1 T 
U iD           (3.43)
   x i 2T x i    P xi 2T x i 

By substituting equation (3.42), the extended continuity equation (3.38) can be written
as:
  U iC  U iD 
  (3.44)
t xi xi

Now, let us introduce a ‘total velocity’, defined as the vector sum of convective and
diffusion velocities, as shown below:
U iT  U iC  U iD (3.45)

By employing this definition of the total velocity, the extended continuity equation
(3.38) can be rewritten as:
 U iT 
 0 (3.46)
t xi
It is interesting that the extended continuity equation (3.46) is similar to the classical
continuity equation (2.1).

Similarly, the extended momentum equations (3.40) can also be rewritten, by


rearranging some of the terms, as:

 U Cj  
 
U iC U Cj  m iDU Cj  m DjU iC  
P   C 2 D C
t x i

x j x i  ij  3  ij m k U k   g j (3.47)
 

Employing the definition of total velocity given in equation (3.45), one can write the
following expression:

25
U iT U Tj  U iC U Cj  U iDU Cj  U DjU iC  U iDU Dj  U iCU Cj  m iDU Cj  m DjU iC  m iD U Dj
(3.48)
Using equation (3.48), equation (3.47) can be rewritten as:

 U Tj  
 T T

U i U j  
P   C
 D D 2
 D C 
 U Dj   (3.49)
  ij  mi U j   ij m k U k    g j
t x i x j xi  3  t

The newly added term m iDU Dj within the total momentum diffusion term has the
product of pressure and temperature derivatives; see equations (3.9) and (3.43), which
were neglected earlier as higher order terms in the Taylor series expansion. Therefore,
one can also drop this term from the right hand side of equation (3.49).

Similarly, one can also rewrite the extended energy equation given in equation (3.41)
in terms of the total velocity as:
 e   U iC e



  
  
T D 
 m i e   P
U iC T
  ij
U Cj
(3.50)
t x i x i  x i  x i x i

By readjusting the total heat diffusion terms, one can obtain the extended energy
equation in terms of the total velocity as given below:

T
  e   U i e 

  
  
T 
  P
U Cj T
  ij
U Cj
(3.51)
t x i x i  x i  x j x i

The extended governing equations given in the total velocity form shown in equations
(3.46), (3.49) and (3.51) may be extremely useful in certain flow configurations as
shown later in this thesis.

3.4.2 Extended total energy equation

In many compressible flow problems, it is preferable to solve the energy equation with
the total energy E as a variable which is defined as:
1 2
 
E  e  U Cj  e  Ek
2
(3.52)

As the first step in obtaining the energy equation in terms of the total energy E, one
needs to derive the mechanical energy equation, which is obtained by multiplying the
momentum equation by the velocity. It is interesting that one can derive two
completely different forms of the extended mechanical energy equation. When the
extended momentum equation (3.40) is multiplied byU Cj , one can obtain:

U C

 U Cj   U  U
C i
C
U Cj   U C P C
U j
 ijT
 g jU Cj (3.53)
j j j
t x i x j x i
where  ijT is given by equation (3.37).

26
By employing the chain rule of differentiation, equation (3.53) can be rewritten as:

 U Cj U Cj   U C
U Cj


 U iC U Cj U Cj   U C
U C
U Cj
 U C P C
U j
 ijT
 g j U Cj
j i j j
t t x i x i x j x i
(3.54)

Equation (3.54) can be further simplified as given below:


 U Cj U Cj   U iC U Cj U Cj   U Cj U Cj  C P
 ijT
  U Cj   U iC   U j  U C
j  g jU Cj
t xi  t xi  x j xi
(3.55)

Equation (3.55) can be rewritten once again by expressing the terms inside the square
brackets on the left-hand side using the chain rule of differentiation as:
 U Cj U Cj   U iC U Cj U Cj  C
  U Cj  C 
 U iC U Cj  C  U i 
C 
 U j  U j  U j 
t x i  t t x i x i 
(3.56)
P  ijT
 U Cj  U Cj  g jU Cj
x j x i
Equation (3.56) is simplified further as shown below:
 U CjU Cj   U iC U CjU Cj    U Cj   U iC U Cj 
C    U iC
C   
 U j   U  
j 
t xi  t xi  t xi  
P  ijT (3.57)
 U Cj  U Cj  g jU Cj
x j xi

It is interesting to note that


  U iC

 
in equation (3.57) is not zero based on the
t xi
extended continuity equation shown in equation (3.38). It is important to observe from
equation (3.38) and (3.46) that the material derivative is zero only when written in terms
of the total velocity. Therefore, equation (3.57) can be expressed as:
  U Cj  U iC U Cj  m iD 

t

 U j  
C 2

xi
C
 C 2 C
U i U j   U j 
 t
 
xi
 U Cj
xi 


C P C
 ijT (3.58)
U j U j  g jU Cj
x j xi
1 C 2
C
Equation (3.58) can be rewritten in terms of local kinetic energy Ek 
2
 
U j , due to
C
convective velocityU j , as:

27

 E kC  
 U iC E kC  1  2
   U Cj  

  1 C C 2
 
t

xi t  2  xi  2 U i U j 

U C

  U Cj  
 U iC U Cj C m
U j
 iD  C P C
 ijT
 g jU Cj (3.59)
j    U j U j
 t x i x i  x j x i

Further, equation (3.59) can be expanded using the chain rule of differentiation as:
 E kC   U iC E kC  1 C  U j  1 U Cj 1 C  U iC U Cj  1
C
C C C
U Cj
  Uj  U j  Uj  U i U j
t xi 2 t 2 t 2 xi 2 xi

U C

 U Cj   U U
C
C
i U Cj   U 
C 2 m iD C P
 U j C
U j
 ijT
 g jU Cj
j j j
t xi xi x j xi
(3.60)

The following equation can be obtained by simplifying equation (3.60):


 E kC   U iC E kC  1 C U j  1 C U i U j 
C C C

  Uj  Uj 
t xi 2 t 2 xi
1  U Cj U Cj  m iD C P
 ijT
C
U j  C
Ui C
 Uj   2
 U j C
U j  g jU Cj (3.61)
2  t xi  xi x j xi

By expanding the terms in square brackets in equation (3.61) using the chain rule of
differentiation, equation (3.61) can be simplified to:

 E kC

 
 U iC E kC 1 C
- Uj
      U   U 
2
C
i C 2 m iD
 U Cj
P
 U Cj
 ijT
 g jU Cj
  j
t x i 2 t x
 i  x i x j x i

(3.62)

Further, employing the extended continuity equation (3.38), equation (3.62) can be
further simplified to obtain the final form of the mechanical energy equation as:

 E kC

 
 U iC E kC
 E kC
m iD  PU j

C
P
U Cj


  ijT U Cj  ijT
U Cj 
 g j U Cj (3.63)

t x i x i x j x j x i x i

The extended mechanical energy equation (3.63) can be added to the extended thermal
energy equation (3.41) to obtain the extended total energy equation as:
 E C   U iC E C  m iD  PU iC    ij U j 
T C
q iT
   Ek    g jU Cj
t xi xi xi xi xi (3.64)
C
where the total energy E is given by:

28
1 C
EC  e 
2
 
Uj
2
 e  EkC
(3.65)

Further, one can also derive the total energy equation based on the total velocity-based
momentum equation (3.49). One can obtain the total mechanical energy equation by
multiplying equation (3.49) by the total velocity U Tj as:

U T

 U Tj U T 
 T T
 T P
U i U j  U j T
U j
~ijT T
U j
 U Dj
 g jU Tj
  (3.66)
j j
t x i x j x i t
where ~T is given by:
ij

2
~ijT   ijC  m iDU Dj   ij m kDU kC (3.67)
3

By employing the chain rule of differentiation, one can rewrite equation (3.66) as:
 U Tj U Tj  T
U Tj  U iT U Tj U Tj  T T
U Tj
 U j   U i U j 
t t xi xi
(3.68)
U
P T T
U j
~ijT T
U j
 U Dj
 g jU Tj
 
j
x j xi t

Equation (3.68) can be simplified further as given below:


 U Tj U Tj   U iT U Tj U Tj  T
 U Tj T
U Tj 
  U j  Ui 
t x i  t x i 

U
P T
U j T
U j T
~ijT
 U Dj
 g jU Tj
  (3.69)
j
x j x i t

Equation (3.69) can be rewritten once again by expressing the terms inside the square
brackets on the left-hand side using the chain rule of differentiation as:
 U Tj U Tj   U iT U Tj U Tj  T
  U Tj  T 
 U iT U Tj  T  U i 
T 
 U j  U j  U j 
t x i  t t x i x i 

 U Tj
P
 U Tj
~ijT
 U Tj
 U Dj 
 g jU Tj
 (3.70)
x j x i t
Equation (3.70) is simplified further as shown below:
 U Tj U Tj   U iT U Tj U Tj    U Tj   U iT U Tj 
T    U iT  
 U j    U Tj   
t xi  t xi  t xi  

 U
P T
U jT
~ijT T
U j
 U Dj
 g jU Tj
  (3.71)
j
x j xi t

29
  U iT
Based on the total continuity equation (3.46), one can write the terms 
  in
t xi
equation (3.71) as zero and hence equation (3.71) can be simplified to:
 U Tj U Tj   U iT U Tj U Tj  T
  U Tj   U iT U Tj 
 U j   
t xi  t xi 

 U
P T T
U j
~ijT T
U j
 U Dj
 g jU Tj
  (3.72)
j
x j xi t

1 T 2
T
By introducing the local total kinetic energy E k 
2
 
U j due to the total velocity U Tj ,
equation (3.72) can be rewritten as:
 E kT   U iT E kT    1 2  1 T 2
    U Tj     U i U j  
T

t xi t  2  xi  2 

 U Tj 

  U Tj  

 U iT U Tj 

U T P
 U T
~ijT
 U T
 U Dj
 g jU Tj
  (3.73)
 j j j
 t xi  x j xi t

One can simplify equation (3.73) to obtain the following equation:


 E kT   U iT E kT  1 T  U j  1 T  U i U j  1
T T T
T
 U Tj T
U Tj 
  Uj  Uj  U j  Ui 
t x i 2 t 2 x i 2  t xi 
(3.74)
 U Tj
P
 U Tj
~ijT
U j T

 U Dj   g U T
j j
x j x i t

By expanding the terms in square brackets in equation (3.74) using the chain rule of
differentiation, equation (3.74) can be simplified to:

 E kT

 
 U iT E kT


 PU Tj
P
 

U Tj  ~ijT U Tj ~ T
  ij
U Tj T
U j
 U Dj 
 g j U Tj
 
t x i x j x j x i x i t
(3.75)

The extended mechanical energy equation (3.75) can be added to the extended thermal
energy equation (3.41) to obtain another form of the extended total energy equation:

30

 E T   U T
i ET    q i


 PU iT
P
U iD  ~ij U j
 
T T
~ T
  ij

U Tj 
t x i x i x i x i x i x i
(3.76)
  ijT
U Cj
 U Tj

 U Dj   g U T
j j
x i t

where E T  e  12 U Tj  is the local total energy based on the total kinetic energy of the
2

fluid. This form of the total energy equation was found to be helpful in solving wall-
bounded flows.

The extended Navier-Stokes equations derived above are reproduced again below as a
summary and these equations are solved to obtain the solutions presented in the
subsequent chapters for gas flow problems in the presence of strong density and
temperature gradients.
Continuity equation:
 U iC  m D
  i (3.38)
t xi xi

Momentum equation:

 U Cj    U i
C
U Cj    P   T
ij
 g j (3.40)
t x i x j x i

Momentum equation in the total velocity form:



 U Tj   T
 T P   C
 D D 2 D C 
 U Dj  
U i U j     ij  m i U j   ij m k U k    g j (3.49)
t x i x j xi  3  t

Thermal energy equation:


 e   U iC e  q iT U iC T
U Cj
  P   ij (3.41)
t xi xi xi xi

Mechanical energy equation:



 E kC

 
 U iC E kC
 Ek

m iD  PU j

C

P
U Cj


  ijT U Cj 
  ijT
U Cj
 g jU Cj (3.63)
 
t x i x i x j x j x i x i

Mechanical energy equation based on the total kinetic energy:



 E kT

 
 U iT E kT

 PU Tj
P 

U Tj  ~ijT U Tj
 
 ~ijT
U Tj
 U Tj

 U Dj
 g j U Tj
  
t x i x j x j x i x i t
(3.75)

31
Total energy equation:
 E C   U iC E C 

q iT
  Ek
m iD  PU iC
 
  
  ij U j T

 g jU Cj
C
 (3.64)
t xi xi xi xi xi

Total energy equation based on the total kinetic energy:


 E T   U iT E T  q i  PU iT  U iD  ~ij U j  ~ T U j
T T T
T
U Cj
   P    ij   ij
t xi xi xi xi xi xi xi
(3.76)
 U Tj
 U D
j   g U T
j j
t

Total molecular momentum transport:


 U Cj U iC  2 U kC 2
T
  
ij
     ij   m iDU Cj  m DjU iC   ij m kDU kC (3.37)
 xi x j  3 x k 3

2
~ijT   ijC  m iDU Dj   ij m kDU kC (3.67)
3

Total molecular transport of heat:


T
q iT    m iD e (3.21)
xi

3.5 Brenner’s extended Navier-Stokes equations

Brenner [2005(a), (b)] also proposed a set of extensions to the classical Navier-Stokes
equations based on his observations of thermophoretic motion. Thermophoresis,
defined as the observed motion of particles against the temperature gradient in an
otherwise quiescent stratified gas, is not completely explained by the CNSE. Brenner
therefore introduced two type of velocities, namely the mass velocity and volume
velocity in his formulation. The velocity appearing in the CNSE, shown in equations
(2.1)–(2.3) and (2.24)–(2.26), was referred to as the mass velocity U im (same as the
convective velocityU iC ). Further, a fictional volume velocity U iV was introduced in the
governing equations. Brenner’s suggestion was to include the volume velocityU iV in the
constitutive relationship of the classical diffusive momentum transport instead of the
mass velocityU im . Hence the constitutive relationship for the molecular momentum
transfer proposed by Brenner is given below:
 U Vj U iV  2 V
 Br
ij        ij  U k (3.77)
 xi x j  3 x k
 

The volume and mass velocities are interrelated by the following expression:
U iV  U im  j iV (3.78)
32
where j iV is defined as:
 ln    1  
jiV       (3.79)
x i   x i 
where  and  are the local thermal diffusivity and density of the gas under
consideration.


Further, by incorporating the definition of  , one can rewrite equation (3.79) as:
C p
  1  
jiV    (3.80)
C p   x i 

In equation (3.80),  is the local thermal conductivity of the fluid. By substituting


equation (3.78) in equation (3.77), the following equation for the molecular transport of
momentum can be obtained:
 U mj U im  2 U km  j Vj j iV  2 V
 Br
ij        ij        ij  j k (3.81)
 xi x  3 x  xi x j  3 x k
 j  k  

Equation (3.81) can be rewritten as:


 j Vj j iV  2 V
 Br
ij  C
ij       ij  j k (3.82)
 xi x j  3 x k
 
where  ijC is the classical molecular transport of momentum terms based on the mass
velocity, as given in equation (2.24).

By incorporating the extended molecular momentum transport terms, given by equation


(3.82), in the classical momentum equation (2.2), the extended momentum equation due
to Brenner was obtained as given below:

 U mj    U i
m
U mj    P 
 ijBr
 g j (3.83)
t x i x j x i

Further, Brenner also modified the classical heat diffusion term as:
j iq  q i  Pj iV (3.84)
where q i is the classical heat diffusion term given in equation (2.26) and P the local
pressure in the flow field. The modified heat diffusion term in equation (3.84) and the
extended momentum diffusion term in equation (3.82) can be incorporated in the
classical energy equation (2.25) in order to obtain the extended energy equation due to
Brenner as:
 e   U im e


j q
 i P
U mj 
  ijBr
U mj
(3.85)
t x i x i x j x i

33
Furthermore, a similar U im to U iV modification for the velocity appearing in the no-slip
tangential velocity boundary condition at solid surfaces was also suggested by
Brenner. However, the normal velocity at the solid wall was retained as the mass
velocity U im . The extensions suggested by Brenner retained the classical continuity
equation given in equation (2.1) without any further modifications.

The extended equations proposed by Brenner are summarized here for ease of
comparison with the set of extended equations derived in this thesis.

Continuity equation:
 U im 
 0 (2.1)
t xi

Momentum equations:

 U mj   U i
m
U mj    P   Br
ij
 g j (3.83)
t x i x j x i

Thermal energy equation:




 e   U im e j q
 i P

U mj
  ijBr
U mj
(3.85)
t x i x i x j x i

Molecular momentum transport:


 U mj U im  2 m  j V V  2 V
 Br
ij        ij  U k    j  j i    ij  j k (3.81)
 xi x j  3 x k  x i x j  3 x k
   

Molecular transport of heat:


j iq  q i  Pj iV (3.84)

3.6 Comparison of the extended Navier-Stokes equations

Another set of extended Navier-Stokes equations was proposed by Oettinger [2005] in


his phenomenological generic theory. He demonstrated that in the generic formulation,
one would obtain the classical Navier-Stokes equations when the terms associated with
mass density in the friction matrix were identically zero. Then, by including non-zero
terms associated with mass density in the friction matrix, a revised set of governing
equations was derived that included two velocities, similar to the volume and mass
velocities proposed by Brenner [2005(a), (b)]. He further argued that the classical
governing equations historically ignored mass diffusivity on the basis that the diffusive
mass flux was assumed to be zero. However, he observed that the associated
momentum and energy fluxes might not be zero. His formulation included the
momentum and energy fluxes, both of which are entropy producing, making the process
34
of mass diffusion irreversible. He felt that the ability of mass diffusion to produce
entropy was missing from the classical Navier-Stokes equations.

Table 3.1 Comparison of the extended equations proposed in the present work and by
Brenner [2005(a),(b)]

Type of Present work Brenner


Equation
  U iC



m iD   U im


0

Continuity t xi xi t xi

 U Cj    U C
i U Cj    P 
 U mj    U i
m
U mj    P
t xi x j t xi x j
Momentum
 ijT  ijBr
  g j   g j
xi xi
 e   U iC e

 q T
 i P
U iC  e  U im e  jiq 
U j m

  P
t xi xi xi t xi xi x j
Energy
T
U Cj U mj
 ij  Br
xi ij
xi

Molecular  1  1 T    1  
transport of m iD       j iV    
  x i 2T x i  C p   x i 
mass
 U Cj U iC  2 U kC  U mj U im  2 m

Molecular  T
ij         ij   Br
ij        ij  U k
 xi x j  3 xk  xi x j  3 x k
transport of   
momentum 2  j Vj jiV  2 V
 m iDU Cj  m DjU iC   ij m kDU kC       ij  j k
3  xi x j  3 xk
 

Molecular
transport of T j iq  q i  Pj iV
T
heat q  
i  m iD CV T
xi

Based on the work of Brenner [2005(a),(b)] and Oettinger [2005], Greenshields and
Reese [2007] predicted shock structures using the extended form of the Navier-Stokes
equations given by Brenner [2005(a)] and obtained better agreements with experimental
measurements of Alsmeyer [1976]. However, they had to include the mass diffusion in
the continuity equation (2.1) proposed by Brenner [2005(a),(b)]. It is interesting that it
was not possible for them to obtain convergence of the solution without the inclusion
of the mass diffusion. It can be easily observed that the inclusion of mass diffusion in
35
the continuity equation to obtain converged solutions was a major deviation from the
work of Brenner [2005(a)]. Further, they also reported that the results obtained with
modified form of the equations suggested by Brenner [2005(a)] (with inclusion of mass
diffusion in the continuity equation) displayed unphysical behaviour when the
coefficient of volume diffusion,  in equation (3.79) exceeded the kinematic viscosity
     . They speculatively claimed that this could be attributed both to instabilities
in temporal disturbances and to a spurious phase velocity-frequency relationship. Later,
Chakraborty and Durst [2007] compared the two independent and apparently dissimilar
considerations proposed by Durst et al [2006] and Brenner [2005(a), (b)].

It was concluded from this study that the constitutive relationships of heat and
momentum transport in ideal gas flows with strong density and temperature gradients,
obtained by these two different proposals, possessed physical and mathematical
equivalences at small gradients. Here, the similarity was achieved by casting the
momentum flux in an equivalent linear form based on a gradient-diffusion hypothesis
under the assumption of constant pressure condition.

Although there have been a number of attempts to compare the two sets of extended
equations, there are conspicuous discrepancies between them:

1. The extended equations proposed by Brenner [2005(a),(b)] and Oettinger [2005]


were obtained based on phenomenological observations, i.e. they were obtained
to explain certain physical or experimental observations. For example, Brenner
[2005(a),(b)] intended to explain the thermophoretic motion and, hence,
adjusted the constitutive relationships to that phenomenon only. However, the
extended equations presented in this thesis were obtained based on the first
principles and are generic in nature, i.e. the formulation and derivations of these
equations were not carried out to explain a particular phenomenon. Therefore,
one can attempt to solve any gas flow problem with density and temperature
gradients with the set of equations presented in this thesis.

2. It is interesting to observe from equation (3.80) and Table 3.1 that the diffusion
velocity is in the direction of the density gradient in the case of Brenner’s
extension whereas it is in the opposite direction to the density and temperature
gradients, as shown in equation (3.7), in the present derivations. As Brenner
attempted to explain the thermophoretic motion, he observed the motion of the
particles to be in the direction of the density gradient and, hence, he expressed
the diffusion velocity to be in the same direction as the density gradient.
However, in general, it is customary to express diffusion transport as functions
of negative gradients of field variables, see equations (2.20), (2.26) and (3.7). It
is argued, as shown in chapter 8, that the incomplete understanding of
thermophoretic motion led to the representation of diffusion velocity as shown
in equation (3.80).

36
3. The diffusion velocity in Brenner’s derivations is a function only of the density
gradient, see equation (3.80), whereas the diffusion velocity is a function of
both density/pressure and temperature gradients in the present derivations, see
equation (3.7). Further, the coefficient of diffusion is also different in both
cases.

4. One more striking discrepancy is the absence of mass diffusion in the continuity
equation in Brenner’s extensions, as shown in equation (2.1). As there is no
apparant diffusion mass transport observed in the thermophoretic motion,
Brenner’s continuity equation did not include any diffusion transport. Further,
it is interesting that since Brenner‘s extended equations were derived to explain
thermophoretic motion, the absence of mass diffusion from the continuity
equation also made the volume velocity fictious. However, the diffusion
transport of mass, shown in equation (3.7), is a physical and real quantity and
also measureable. It is also further argued that the presence of mass diffusion
leads to the inclusion of extended constitutive relationships for momentum and
heat, as shown in equations (3.37) and (3.21), respectively, in the present
derivations.

5. Furthermore, the constitutive relationships of diffusion transport of momentum


and heat in both cases are different, as shown in equations (3.37) and (3.81) and
equations (3.21) and (3.84), respectively.

The above differences suggest that the extended form of the Navier-Stokes equations
derived in this thesis is completely different from those suggested by Brenner
[2005(a),(b)] and Oettinger [2005]. It is argued that the extended equations derived in
this chapter have a strong physical basis and, hence, can be employed in solving gas
flows with strong temperature and density gradients in order to obtain accurate
predictions. A number of such flow problems are solved in subsequent chapters in
order to substantiate this argument.

37
38
Chapter 4

NUMERICAL PREDICTIONS OF GAS FLOWS


THROUGH MICROCHANNELS
4.1 Introduction and Literature Review

The entire field of flows through microchannels has become a very popular research
area due to an increasing interest in Micro-Electro-Mechanical Systems (MEMS), as
well as bio-mechanical, Lab-on-the-Chip Systems (LCS); see Karniadakis et al
(2005). These flows remain a major interest to fluid mechanics research since the very
first experimental investigations; see Arkilic et al [2001] and Maurer et al [2003]. The
MEMS systems typically employ small channels with characteristic dimensions in the
range of 1-100µm; see Gad-el-Hak [2002]. Analysis of flow of gases through straight
channels of such dimensions is an integral part of designing MEMS systems. A lot of
literature is available on the gas flow experiments conducted in microchannels; see
Pfahler et al (1991), Pong et al (1994), Harley et al (1995), Liu et al (1995) and Shih
et al (1995, 1996). It can be observed from these literatures that under some inlet and
outlet pressure conditions, the experimentally measured mass flow rates of gases were
higher than those computed/calculated employing the classical Navier-Stokes
equations (CNSE). This vital observation had puzzled and excited researchers all over
the world over the last few decades and a large number of publications on this topic
have emerged.

In the absence of any other sound physics based answers to the lower mass flow rates
obtained with CNSE, employing the no-slip boundary condition at the solid wall, in
comparison with the corresponding experimentally measured values, the theoretical
treatments in the published literature suggested employing a slip velocity at the wall;
see Arkilic et al [2001] and Maurer et al [2003]. Further, this velocity was assumed to
be the slip velocity proposed by Maxwell (1879) who suggested that the velocity at a
molecularly smooth solid wall can be calculated by:
C
 2    dU 1
US  l (4.1)
   dx 2

In equation (4.1), l is the molecular mean free path, U1C the tangential velocity
component and x2 the normal direction to the wall. Further,  is the tangential
momentum accommodation coefficient. The expression for the slip velocity shown in
equation (4.1) is based on the assumption that one fraction  of the molecules
interacting with the wall is reflected in a diffusive way and the rest is reflected
specularly. Thus, by assuming that a certain fraction of molecules retain their
tangential momentum after bombarding the solid wall, it was possible to obtain a non-
zero wall velocity with the help of equation (4.1). However, it is important to note
that the fraction  had to be experimentally determined, i.e. the theoretical predictions
had to be empirically fitted to the experimental measurements in gas flow treatments
based on slip-velocity in micro-conduits. So far, the successful theoretical and
numerical treatments of microchannel flows were based on the CNSE, shown in
equations (2.1)–(2.3) and (2.24), subjected to the Maxwell slip velocity boundary
condition at the solid walls, shown in equation (4.1); see Arkilic et al (1994) and
Arkilic and Schmidt [1997].

As shown by Arkilic and Schmidt [1997], it is possible to obtain an analytical solution


for the mass flow rate of an isothermal ideal gas flow through two-dimensional
microchannels with the slip boundary condition at the walls as:
2
H 3 wPo2  Pin    Pin   
m       1  12 Kn    1  (4.2)
24LT  Po     Po   
  
where H, L and w are the height, length and width of the considered microchannel,
respectively. Further, Pin and Po are the inlet and outlet static pressures, respectively
and T the absolute static temperature of the fluid in the channel.  is the dynamic
viscosity of the fluid and  the specific gas constant. Kn is the Knudsen number and
2 
 is given by   where  is the tangential momentum accommodation

coefficient; see Arkilic and Schmidt [1997]. It is well-known that if  becomes zero
in equation (4.2), one obtains the conventional no-slip boundary condition for the flow
velocity and hence, one can obtain the corresponding no-slip mass flow rate through
the microchannel as:
2
H 3 wPo2  P  
m   in   1 (4.3)
24 LT  Po 
 

Though the research community has widely accepted the ‘slip-velocity’ based models
for predicting gas flows through microchannels, it can be proven easily that employing
the Maxwell slip-velocity, valid only for the case of molecularly smooth walls, in
microchannel flows is highly objectionable. Typical wall roughness values in
microchannels are of the order of 10 8 - 10 7 m whereas the diameter of molecules is of
the order of 10  10 m. As shown by Mo and Rosenberger [1990], the apparently smaller
wall roughness values of microchannels are at least 2-3 orders greater than the
molecular diameter and hence, this wall roughness values are good enough to enforce
the no-slip boundary condition to the flow. Therefore, the slip-velocity assumption in
analyzing gas flows through microchannels is questionable and needs to be scrutinized
thoroughly.

When the research work leading to this thesis was underway, it was proposed that the
presence of self-diffusion of mass, driven by strong density/pressure gradients, in gas
flows through microchannels could be the reason for the increased mass flow rates
observed in the experiments. It is interesting to note that since the height of
40
microchannels is very small in comparison to the length, i.e. h  L , the streamwise
pressure gradients are in general, very high and the inlet and outlet pressures can also
be less than the atmospheric pressure under typical operating conditions of the
channel. Hence, as observed in equation (3.9), the self-diffusion caused mass
transport in gas flows through microchannels can be significant and needs to be
incorporated in the theoretical or numerical analysis. With this intuitive supposition, it
was decided to employ the total velocity form of the ENSE, presented in equations
(3.46), (3.49) and (3.67) to numerically analyse flow of gases through microchannels.

Interestingly, as shown later in this chapter, the numerical results obtained with these
equations without any empirical tuning parameters agree excellently with the
experimental measurements. Further, it is argued that the approach presented here
accurately handles the physics behind gas flows through microchannels. It is also
interesting that all the salient characteristics of microchannel flows are also explained
with great insight based on this analysis, thus proving convincingly that the observed
additional mass flow rate in experiments is nothing but the self-diffusion mass
transport driven by the pressure/density gradients. Further, because of the small
dimensions of micro-conduits, it is known that the Knudsen number Kn, defined as the
ratio of the molecular mean free path l to the characteristic length Lc becomes a key
parameter to determine the characteristics of these flows. Conventionally, the height of
the channel is taken as the characteristic length of flows through microchannels and
the Knudsen number is defined as:
l
Kn  (4.4)
H

However, it is also possible to define the characteristic length based on the gradients
of macroscopic quantities such as density or pressure, and the Knudsen number can
also be expressed as:
 1    1 P 
Kn  l    l   (4.5)
  x1   P x1

Equation (4.5) is considered as a more precise choice for the characteristic length; see
Gad-el-Hak [1999]. Interestingly, the length scale for the diffusive mass transport of
isothermal gas flows, shown in equation (3.9), is similar to the one shown in equation
(4.5). Hence, it is important to deduce which definition of Knudsen number provides
a clear description of the flow characteristics in micro-conduits.

4.2 Analytical solutions based on the CNSE with the no-slip boundary
condition

While obtaining the analytical expressions shown in equations (4.2) and (4.3), the gas
flow through the channel was assumed to be fully-developed and the nonlinear
acceleration terms were neglected by Arkilic and Schmidt [1997]. It is generally
assumed that the low Reynolds number creeping flows in microchannels behave like

41
incompressible flows; see Zhang et al [2005]. However, it is easy to observe that the
acceleration terms in the left-hand side of the momentum equations, presented in
equation (2.2), are non-zero because of the variation of density and the velocity profile
does not become fully-developed. Therefore, under certain operating conditions, the
nonlinear terms may assume significance and cannot be neglected. Hence it is
required to estimate the relative influence of the acceleration terms in gas flows
through micro-conduits and neglecting of inertial terms needs to be justified.
Therefore, an analytical solution procedure based on the integral form of the CNSE
with the no-slip boundary condition was developed and the results were compared
with equation (4.3) and experimental measurements of straight microchannels and
capillaries.

x2

x1

x2 (a)

x1

x2 (b)

x1

(c)

Figure 4.1 Schematic representation of the coupling of density and velocity profiles in
gas flows through micro-conduits; Profiles of (a) density (b) streamwise velocity and
(c) mass flow rate
42
The schematic representation of the density and streamwise velocity profiles of gas
flows through microchannels are shown in Figure 4.1. The pressure-gradient driven
gas flow through the channel has higher density at the inlet than that at the outlet
which results in a gradual increase of the streamwise velocity towards the outlet.
Because of large aspect ratios, the development length is negligibly small in
comparison to length of microchannels and hence the cross-stream velocity is zero
everywhere. Therefore, one can easily deduce that the mass flow rate profiles do not
vary along the streamwise direction, i.e. the flow is fully-developed in terms of the
mass flow rate and not in terms of the streamwise velocity. The geometry and
boundary conditions employed are shown in Figure 4.2. In this figure, h is half-height
of the channel and R the radius of the capillary. U m is the maximum velocity at any
cross-section and Pi and Po are the static pressures at the inlet and outlet boundaries,
respectively.

4.2.1 Derivations for microchannel flows

The governing equations for isothermal gas flows through a microchannel are given
below.

Continuity equation:

 U1C
0
 (4.6)
x1

Momentum equation:
 U 1C U 1C dP 
  U 1C 
(4.7)
   
x1 dx1 x 2  x 2 

In equation (4.7), the streamwise momentum diffusion is neglected in order to simplify


the analytical solution procedure. Since the cross-stream velocity is zero throughout the
domain, the mass flow rate at any given cross-stream location x2 is constant, i.e.
 x1 U 1C x1 , x 2  is constant. It is also important to note that the temperature can be
x2

assumed to be constant for low Mach number flows through micro-conduits. Therefore,
the pressure and hence, the density are functions of only the streamwise coordinate x1 .
To obtain integral solutions, it is customary to employ the profile method, see
Schlichting [1979], Mills [1992] and von Karman [1921], by assuming a polynomial
function for the mass flow rate through the channel as:
U 1C   A  B x2   C  x2  2 (4.8)
U mC   h   h 
subjected to the boundary conditions given below:
At x2  0 ;  U 1C    U mC  (4.9a)

At x2  0 ; 
x 2

U 1C  0 (4.9b)

43

At x2  h ; U 1C  0  (4.9c)

x2
Wall,

Outlet
Inlet h
=const
=const Symmetry, ,
x1

L
(a)
r
Wall,

Outlet
Inlet R =const
=const Axis, ,
x1

L
(b)

Figure 4.2 Geometry and boundary conditions employed in the analytical solutions
based on the integral form of the equations: (a) microchannel and (b) capillary

In equation (4.9a), U mC is the maximum value of the velocity at any given streamwise
location x1 . By applying the boundary conditions given in equation (4.9), it is possible
to evaluate the coefficients in equation (4.8) and the mass flow rate profile can be
expressed as:
U1C   1   x2  2 (4.10)
U mC   h 
Further, the momentum equation, given in equation (4.7), can be expressed in the
integral form as:
h
 U 1C U 1C  h
dP h
  U 1C 
 dx2    dx 2    dx2 (4.11)
x1 dx 0 x 2  x 

0
  01    2  

Term - I Term - II Term - III

The three-terms given in equation (4.11) can be integrated individually as shown below:

Term – I = 

 U 1C U 1C
h
dx 2 =

d h  U 1C   2

 dx2 (4.12)
x1

dx1 0  
0
 

44
By introducing the mass flow rate profile given by equation (4.10) in equation (4.12),
Term-I can be written as:
 2 2
d h  1   x2   
Term – I = U C 2
m   1     dx 2 (4.13)
dx1 0     h   
 
Since the density  is only a function of the streamwise coordinate x1 , it is possible to
rewrite equation (4.13) as:
2 2
d  1  h   x 2  
Term – I = U C 2
    1     dx 2 (4.14)
dx1    0   h  
m

After performing the integration, Term-I becomes:

8 1 dP
Term – I =   2

U mC hT 2 (4.15)
15 P dx1
where  is the specific gas constant.

Similarly, the other terms in equation (4.11) can be integrated as:


dP
Term – II =  h (4.16)
dx1
2 T
Term – III = 
hP
U mC   (4.17)

By substituting equations (4.15)-(4.17), equation (4.11) can be expressed as:


8 1 dP dP 2T
 U mC  hT 2 U mC   0
2
h  (4.18)
15 P dx1 dx1 hP

After separating the variables, it is possible to integrate equation (4.18) in x1 to obtain


the following form:
P      
16h 2 T ln in  U mC  60TL U mC  15h 2 Pin2  Po2  0
2
(4.19)
 Po 

The quadratic equation in  U mC  can easily be solved to obtain the peak value of the
mass flow rate profile as given below:
 60 TL  60 TL 2  960h 4 T ln  Pin P 
 P2  P2
 in o 
C  o 
U 
m (4.20)
P
32 h 2 T ln  in 
 Po 

It is well-known that in fully-developed channel flows, the ratio of average to maximum


mass flow rates is given by:
2
U 1C  U mC  (4.21)
3
45
Further, the mass flow rate through the channel is given by:
  
m  2hw U1C  Hw U1C  (4.22)
where H and w are the height and width of the channel, respectively. To compare the
mass flow rate with the one provided by Arkilic and Schmidt [1997] as given in
equation (4.3), equation (4.11) can be solved neglecting Term – I, to obtain the
following expression for the peak mass flow rate:
C h 2 Pin2  Po2 
U m  (4.23)
4LT
It can be observed that the identical expression of the average mass flow rate is obtained
to the one shown in equation (4.3), given by Arkilic and Schmidt [1997], by employing
the peak mass flow rate, given by equation (4.23). Further, when comparing the results
obtained with equations (4.22) and (4.3), it can be concluded that there are no
significant differences in terms of mass flow rates for the operating pressure values
provided by Arkilic et al [2001] and the inclusion of convective terms did not change
the results in any appreciable manner. Hence, it can be concluded that the nonlinear
convective acceleration terms can be neglected from the theoretical or numerical
analysis of isothermal gas flows through straight microchannels considered in this
study. However, for larger channels with higher Reynolds number gas flows, the
acceleration terms may have to be incorporated and the result presented in equation
(4.20) can be used in such cases. It is necessary to stress here that this important
conclusion helps in simplifying the numerical solutions described in this chapter and
ensuring the analytical solutions feasible with ENSE, as described in the next chapter.

4.2.2 Derivations for capillary flows

The geometry and boundary conditions employed to obtain the integral solutions for
gas flows through capillaries are shown in Figure 4.2(b). The CNSE for the case of
isothermal axisymmetric compressible ideal gas flows are given below.
U1C 
Continuity equation: 0 (4.24)
x1

Momentum equation:

 U 1CU 1C 
dP

1   U 1C 
(4.25)
 r 
x1 dx1 r r  r 
The profile method explained in section 4.2.1 resulted in a mass flow rate profile
similar to the one shown in equation (4.10) and hence, the following mass flow rate
profile was employed in the analysis:
U 1C   1   r  2 (4.26)
U mC   R 
subjected to the boundary conditions given below:
At r  0 ;  U 1C    U mC  (4.27a)

46

At r  0 ;
r
U1C  0   (4.27b)

At x2  R ; U1C  0   (4.27c)

The integral form of the momentum equation shown in equation (4.25) can be written as
given below:
R
 U 1C U 1C  R
dP R
1   U 1C 
 dr    dr    r dr (4.28)
x1 0 dx1 r r r 

0
     0 
Term  I Term _ II Term _ III

On integrating, equation (4.28) results in the following form:


8 1 dP dP 4T
 U mC  RT 2 U mC   0
2
R  (4.29)
15 P dx1 dx1 RP

After separating the variables and integrating, the peak value of the mass flow rate
profile is given by:
 120 TL  120 TL 2  960 R 4 T ln  Pin P 
 P2  P2
 in o 
C  o 
U 
m (4.30)
P
32 R T ln  in 
2

 Po 

It is well-known in fully-developed pipe flows that the ratio of the average to the
maximum mass flow rate is given by:
1
U 1C  U mC  (4.31)
2

Further, the mass flow rate through the capillary can be obtained by:
m  R 2 U1C (4.32)

Furthermore, when equation (4.28) was solved by employing terms II and III only, the
expression for the peak value of the mass flow rate profile could be obtained as:
R 2 Pin2  Po2 
U mC  (4.33)
8LT
Therefore, the average mass flow rate through the capillary can be expressed as:

U 1C 

R 2 Pin2  Po2


D 2 Pin2  Po2   (4.34)
16LT 64LT
where D is the diameter of the capillary considered.

47
4.2.3 Importance of diffusion mass transport

As shown above, one can express the mass flow rates of ideal gas flows through
straight microchannels and capillaries employing equations (4.22) and (4.32),
respectively. It is well known that under certain operating conditions, the mass flow
rates obtained with the no-slip boundary condition at the solid wall deviate from the
experimentally measured values, see Maurer et al [2003], Arkilic et al [2001] and
Gad-el-Hak [1999, 2002], which led to the introduction of the ‘slip-flow’ theory as
described in section 4.1. As shown in the subsequent section, the ENSE were
employed to predict the characteristics of gas flows through microchannels in this
thesis. Before embarking on the numerical simulations with the ENSE, it was required
to establish the flow conditions in which the additional mass flow rate due to self-
diffusion needed to be considered in the analysis.

x2

x1
H
Diffusive Convective
flow flow

Figure 4.3 Convective and diffusive parts of the velocity profile in microchannel flows

The additional diffusion mass flow rate in a straight microchannel can be computed
employing equation (3.9) as given below:
 1 dP 
m 1D  U 1D H   
 D  H (4.35)
 P dx1 

For the case of microchannel flows, the ratio of mass flow rates due to diffusion and
convection can be obtained employing equations (4.35) and (4.3) as:
m D 12 2
  C  (4.36)
m PH 2

Employing the following well-known relationship for the Knudsen number Kn:
 Ma l
Kn   (4.37)
2 Re H

48
and employing the definitions of the Mach and Reynolds numbers given by
UC  HU 1C 
Ma  1
C
Re  
   , respectively, one can derive the expression for the mean
molecular free path l , as:
 
l (4.38)
2 C
where C is the local velocity of sound and the kinematic viscosity of the fluid. Using
the expression for the mean molecular free path l given in equation (4.38), it is
possible to deduce the ratio of diffusion and convective mass flow rates  for
microchannel flows from equation (4.36) as:
m D 24
  C  Kn2 (4.39)
m 

Similarly, the ratio of diffusion and convective mass flow rates for gas flows through
capillaries can be expressed, employing equations (3.9), (4.32), (4.34) and (4.36), as:
m D 64
  C  Kn 2 (4.40)
m 
where the diameter of capillary D is used as a characteristic dimension in defining the
Knudsen number Kn.

Employing equations (4.39) and (4.40), one can evaluate the Knudsen number at
which the self-diffusion mass transport assumes significance over the convective
transport. Table 4.1 provides values of the ratio of mass flow rates  at various
Knudsen numbers for ideal gas flows through microchannels and capillaries. One can
clearly observe that the error of around 5% in mass flow rate will be present if the self-
diffusion mass transport is not considered in the analysis, even at a relatively smaller
Knudsen number of 0.05. Further, the influence of self-diffusion increases rapidly
beyond a Knudsen number value of 0.25. It is also interesting to note in Table 4.1 that
the influence of self-diffusion assumes significance at even smaller Knudsen numbers
in capillary flows than flows through microchannels.

Table 4.1 Ratio of mass flow rates  as a function of Knudsen number

 0.01 0.05 0.1 0.25 0.5 0.75 1 2 5 10


Kn
Channel 0.036 0.081 0.114 0.181 0.256 0.313 0.362 0.512 0.809 1.144
flows
Kn
Capillary 0.022 0.05 0.07 0.111 0.157 0.192 0.222 0.313 0.495 0.701
flows

49
In order to validate the important result shown in Table 4.1, the convective mass flow
rates obtained with equations (4.3) and (4.22) are compared with the experimental
measurements of Maurer et al [2003] in Figure 4.4. As mentioned earlier, it was
found that the mass flow rates predicted by equations (4.3) and (4.22) were near
identical and hence only one set of values is shown. The Knudsen number was
evaluated at the average of the inlet and outlet pressures. It is clearly evident from
Figure 4.4 that the theoretical estimate of the influence of diffusion mass flow rate
shown in equation (4.39) and Table 4.1 is accurate and one can clearly observe the
deviation of the experimental measurements from the values predicted by the CNSE
employing the no-slip boundary condition beyond a Knudsen number of 0.1. Hence,
one is encouraged to take the self-diffusion driven mass transport into account in
predicting the characteristics of gas flows through straight microchannels and
capillaries employing the ENSE.

CNSE
Expt._Maurer et al [2003]

Figure 4.4 Deviation of mass flow rate predicted by the CNSE employing the no-slip
boundary condition at the wall from the experimental measurements of
Maurer et al [2003]

4.3 Governing equations, geometry and boundary conditions

The total velocity based ENSE presented below, simplified from equations (3.46),
(3.49) and (3.9), were employed to solve steady isothermal gas flows through straight
microchannels.
U iT 
Continuity equation: 0 (4.41)
xi

50
~T
 P  ij
Momentum equations:
x i
T T

U i U j  
x j
x i
(4.42)

Molecular momentum transport:


C
 U j U iC  2 C
~ijT         ij  U k  m iDU Dj  2  ij m kDU kC
 xi x j  3 x k 3
 
[4.43, same as equation (3.67)]
 1    1 P 
Self-diffusive transport of mass: m iD           (4.44)
  xi   P xi 

It is well-known that the available computational fluid dynamics (CFD) codes have
been developed for the CNSE employed with the no-slip boundary conditions,
presented in equations (2.1)-(2.3). The computations mentioned here were carried out
in the commercially available CFD software, Fluent 6.3. The additional terms in the
ENSE were incorporated into the solver as mass and momentum source terms
employing User Defined Functions (UDF). The UDF code is provided in Appendix.

x2
Wall, ,

Outlet
Inlet h
Symmetry,
x1

L
(a)
x2
Wall, ,

Outlet
Inlet
h
Symmetry,

x1
L
(b)

Figure 4.5 Geometry and boundary conditions employed in the numerical simulations
of straight two-dimensional microchannels: (a) CNSE and (b) ENSE

The two-dimensional geometry and boundary conditions employed in the simulations


of straight microchannels employing both the CNSE and the ENSE are presented in
Figure 4.5. The published experimental measurements of Maurer et al [2003] were
51
used to compare the numerical results obtained for the straight microchannels and the
salient experimental conditions are summarized in Table 4.2. The grid counts in the
streamwise and cross stream directions were 30 and 5000, respectively. At the inlet
and outlet boundaries, constant static pressure values estimated based on the
experimental conditions shown in Table 4.2 were specified. Symmetry boundary
condition was employed at the middle plane of the channel since only half of the
channel was simulated. Second order discretization schemes were employed for all
variables whereas the PISO (Pressure Implicit Splitting of Operators) scheme was
used for pressure-velocity coupling.

Table 4.2 Summary of experimental conditions employed by Maurer et al [2003]

Experimental parameters Maurer et al [2003]

Reported channel height, H μm  1.14

Width, w μm  200


Length, L mm 10
Gas used Helium
Gas constant [J/Kg.K] 2077
Temperature, [K] 293
Viscosity of the gas [10-5 Pa s] 1.99
Inlet Pressure range [kPa] 26-507
Outlet Pressure [kPa] 12-101
Outlet Knudsen number Kno  1.509-0.179

It is necessary to explain the wall boundary condition employed in the simulations


with the ENSE. It is a well-known practice to employ the no-slip boundary condition
at the solid wall in the classical fluid mechanics as shown in Figure 4.5(a).
Interestingly, based on experimentally observed higher mass flow rates in comparison
to those obtained with the classical no-slip boundary condition, the slip-flow theory
was introduced in numerical simulations of gas flows through micro-conduits where
the slip-coefficient needed to be experimentally determined. It is shown clearly by Mo
and Rosenberger [1990] that unless the wall is atomically smooth, the no-slip
boundary condition for the velocity is valid at all the engineering surfaces.

To understand the nature of molecules-wall interactions, one can refer to the


schematics shown in Figure 4.6. As shown in Figure 4.6(a), unlike the molecules
moving in far-away distances from the solid wall, the molecules caught in between the
52
roughness elements are expected to lose their momentum due to multiple
bombardments onto the wall roughness elements and hence the molecules tend to
attain the velocity of the wall. The roughness elements are a minimum of two orders
of magnitude higher than the diameter of molecules in a typical microchannel; see
Arkilic et al [2001]. Therefore, as shown schematically in Figure 4.6(b)-(d), the
convection, diffusion and total velocity profiles are expected to satisfy the no-slip
boundary condition at the wall.

Molecule

Mean
roughness
height

Wall Wall

(a) (b)

Mean
roughness
height

Wall Wall

(c) (d)

Figure 4.6 Schematic representation of flow characteristics near the solid wall in gas
flows through microchannels (a) Effect of wall roughness; (b) Convective velocity
profile; (c) Diffusion velocity profile; (d) Total velocity profile

However, unlike the convection velocity profile, depicted in Figure 4.6(b), which
gradually increases from a value of zero at the mean height of roughness elements at
the wall to the maximum at the mid-section of the channel, the diffusion velocity
depicts a sharp jump immediately beyond the roughness elements of the wall. Since
gas flows through straight microchannels behave like fully-developed flows, the
pressure remains constant along the cross-stream direction and hence, the pressure
gradient values at the wall and mid-section of the channel are identical. Therefore,
from equation (4.44), one can observe that the magnitude of diffusion velocity is the
53
same in the cross-stream direction at any given streamwise location. It is easy to
conclude that when the boundary conditions are employed at the mid-height of
roughness elements of the solid wall, the total velocity profile, defined as the sum of
convective and diffusive velocity profiles, displays a slip-like behaviour at the solid
wall as shown in Figure 4.6(d).

It is important to note that the physical reason for this ‘slip-like’ non-zero velocity at
the wall, i.e. at the mean height of roughness elements, is the self-diffusion mass
transport driven by pressure/density gradients and not the tangential momentum
accommodation based slip concept adopted by the ‘slip-flow’ models proposed in the
microchannel literature like Arkilic et al [2001] and Maurer et al [2003]. Based on the
above arguments, the boundary conditions at the wall were introduced in the
simulations as shown in Figure 4.5(b). Unlike the convection velocity, the normal
gradient of the diffusion velocity at the solid wall is zero in straight micro-conduits,
i.e. the self-diffusion caused mass flow results in insignificant momentum transfer to
the solid wall or the diffusion flow experiences negligible resistance from the solid
wall. This important observation helps in providing physically sound explanations to
some of the apparently puzzling characteristics of micro-conduit flows such as the
‘Knudsen paradox’, as described later in this chapter.

4.4 Results and discussions

4.4.1 Velocity profiles

Since the Knudsen number increases along the length of the channel, the relative
significance of the self-diffusion mass transport with respect to the convective mass
transport can also vary significantly as shown in equation (4.39). Therefore, the flow
characteristics can be governed purely by convection in some regions of the channel
whereas the self-diffusion can assume huge significance at some other regions of the
channel. As examples, two velocity profiles are discussed here to explain the
changing dynamics at different sections of the channel.

In Figures 4.7 and 4.8, the total velocity profiles are presented at two different axial
locations and pressure ratios. The convection and diffusion velocity components of the
total velocity are also shown in these figures. Further, the velocity profiles obtained
numerically by employing the CNSE with the no-slip boundary condition are also
provided for comparison. In Figure 4.7(a), at half of the length of the channel, i.e.
x1  0.005m, one can observe that there is almost no difference between the solutions
of the CNSE and the ENSE and the diffusion velocity is negligibly small. The local
value of pressure at this location was found to be 4.9 bar which resulted in the local
Knudsen number of 0.035. One can observe from equation (4.39) that the diffusion
mass flow rate is negligibly small at this Knudsen number and hence, the velocity
profiles obtained with the CNSE and the ENSE are near-identical. On the contrary, in
Figure 4.7(b), it can be observed that there exists a significant deviation in the velocity
54
profiles at the outlet plane of the channel. The convective velocity component
obtained with the ENSE is much less than that calculated with the CNSE.

6.E-07
Convection_component
Diffusion_component
5.E-07 Total_velocity
Classical_conv_velocity
Cross-stream coordinate, m

4.E-07

3.E-07

2.E-07

1.E-07

0.E+00
0 0.1 0.2 0.3 0.4 0.5 0.6
Axial velocity, m/s
(a)
6.E-07

5.E-07
Cross-stream coordinate, m

4.E-07

3.E-07

2.E-07
Convection_component
Diffusion_component
Total_velocity
1.E-07
Classical_conv_velocity

0.E+00
0 2 4 6 8 10 12 14 16 18 20
Axial velocity, m/s

(b)
Figure 4.7 Comparison of streamwise velocity profiles obtained with the CNSE and
the ENSE for Helium gas flow through microchannel, Pin  7 bar; Po  0.12 bar;
Streamwise locations (a) 0.005m and (b) 0.01m
55
6.E-07
Convection_component
Diffusion_component
5.E-07 Total_velocity
Classical_conv._velocity
Cross-stream coordinate, m

4.E-07

3.E-07

2.E-07

1.E-07

0.E+00
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09
Axial velocity, m/s

(a)
6.E-07
Convection_component
Diffusion_component
5.E-07 Total_velocity
Classical_conv._velocity
Cross-stream coordinate, m

4.E-07

3.E-07

2.E-07

1.E-07

0.E+00
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18
Axial velocity, m/s
(b)

Figure 4.8 Comparison of streamwise velocity profiles obtained with the CNSE and
the ENSE of Helium flow through microchannel, Pin  0.4 bar; Po  0.12 bar;
Streamwise locations (a) 0.005m and (b) 0.01m

56
Further, it can also be observed that the maximum value of the velocity profile at the
centre of the channel obtained with the CNSE is slightly more than the total velocity
obtained with the ENSE. However, it is evident that the average velocity obtained
with the ENSE is higher than that of the CNSE, similar to the observations in the
experiments. It is also interesting to notice in Figure 4.7(b) that the diffusion velocity
is much higher than its convection counterpart since the local Knudsen number at the
outlet is 1.42.

It is evident from the above discussions that the self-diffusion mass transport need not
influence the characteristics of the channel at all sections for the total mass flow rate
through the channel to increase. When the local Knudsen number remains low, the
convective transport dominates the diffusion and determines the flow rate through the
channel. However, when the Knudsen number assumes a relatively higher value, the
diffusion begins to influence the characteristics of the flow. This reveals a very
intimate coupling between the convection and diffusion mechanisms governing the
dynamics of gas flows through microchannels.

Unlike the case shown in Figure 4.7, it is also possible that the self-diffusion transport
plays a dominant role throughout the length of the channel as shown in Figure 4.8.
For the case considered here, the local Knudsen number was found to be 0.43, 0.71
and 1.42 at the entrance, middle and outlet sections of the channel, respectively. One
can immediately deduce from equation (4.39) that the mass transport due to self-
diffusion is expected to be much higher than that of the convection at these Knudsen
numbers. As evidently seen in Figure 4.8, the CNSE with the no-slip boundary
condition predict very less mass flow rate through the channel in comparison to the
solutions obtained with the ENSE. Further, one can also observe that the total
velocity, dominated by the self-diffusion mass transport, is more than three times than
the convective velocity predicted by the CNSE.

4.4.2 Pressure profiles

From the discussions presented above, one can readily infer that the combined
influence of self-diffusion and convective mass flow rates will change the pressure
profiles along the length of the channel. In Figure 4.9, the dimensionless pressure
profiles, defined as P  P  Po  Pin  Po  , are plotted as a function of the streamwise
coordinate for different pressure ratios. It is important to note that the outlet pressure
was kept at a constant value of 0.12 bar, resulting in an exit Knudsen number of 1.42
at all the simulations. The dimensionless pressure profiles obtained employing the
CNSE with the no-slip boundary condition are also plotted in this figure for
comparison.

57
(a) (b)

(c) (d)
ENSE CNSE

Figure 4.9 Comparison of dimensionless pressure profiles along the length of the
channel; Po = 0.12 bar; PR  Pin Po (a) 58.3; (b) 10; (c) 3.33; (d) 1.67

As the first observation, in Figure 4.9, one can clearly see that the pressure profiles
obtained with the CNSE are convex in nature for all the pressure ratios. On the
contrary, the pressure profiles obtained with the ENSE show a gradual variation from
convex to concave as the pressure ratio decreases. At high pressure ratios, the mass
flow rate is governed by the convective transport in most parts of the channel and
hence, the difference between the pressure profiles obtained by the CNSE and the
ENSE is negligible, as observed in Figure 4.9(a). As the pressure ratio is gradually
decreased, the self-diffusion mass transport starts assuming significance over
convection and influences the flow characteristics. Since there is no cross-stream
gradient of the diffusion velocity, the diffusion transport of mass does not cause any
momentum loss to the wall and hence, the pressure drop through the channel becomes
negligible. Therefore, the prescribed pressure difference across the channel is used to
enhance only the convective transport through the channel.

One may say that the convex shape of the pressure profile is prevalent in the
convection dominated region of the channel and a concave profile is obtained when

58
the diffusion transport is dominant. This combined influence of self-diffusion and
convective mass transports alters the curvature of the pressure profile from convex to
concave, as seen in Figure 4.9(b). As shown in the encircled region of Figure 4.9(b),
the convex profile near the convection dominated inlet region slowly gets transformed
into the diffusion dominated concave profile near the outlet. Detailed analysis on the
nature of pressure and pressure gradient profiles are given in the next chapter. When
the pressure ratio is reduced further, the mass transport due to self-diffusion dominates
the convective transport throughout the channel and the resistance for the flow in the
channel goes down significantly. Here, one obtains the complete concave shape of the
pressure profiles, as shown in Figure 4.9 (c) and (d). These interesting characteristics
in microchannel flows are not predicted by the CNSE.

4.4.3 The ‘Knudsen Paradox’

In his experimental study of rarefied gas flows, Knudsen [1909] discovered that as the
pressure was reduced to very low values, the conductance of the gas did not get
reduced continually. On the contrary, the conductance through the channel reached a
minimum and then started increasing as the pressure was further reduced. Since it was
not possible to explain this phenomenon by the CNSE, this is referred to as the
‘Knudsen paradox’. In other words, the ‘Knudsen paradox’ is defined as the
occurrence of a minimum in the conductance of the channel before rising again when
plotted against the Knudsen number. Both the CNSE and the first order slip models
based on the ‘Maxwellian-slip’ theory are unable to predict the ‘Knudsen paradox’ at
all whereas the predictions based on the second order slip models are not accurate
enough; see Zhang [2005]. Hence, it is generally believed that one needs to employ
the Lattice Boltzmann Equation (LBE) based methods or the Direct Simulation Monte
Carlo (DSMC) techniques to predict the ‘Knudsen paradox’ phenomena in
microchannels. Needless to mention, these techniques also involve a number of
assumptions, for example the parameters for molecule-wall interactions etc, which
need to be experimentally verified and determined and hence, there is certain
empiricism involved in these microscopic models, as well.

The microchannel results obtained with the ENSE were employed to predict the
‘Knudsen paradox’. As it has already been shown in this chapter, there are no
underlying assumptions either in the derivations or in the simulations in the case of the
ENSE. In Figure 4.10, the conductance of the microchannel predicted by the ENSE,
defined as Cˆ  M T Pin  Po  , is plotted against the average Knudsen number,
evaluated at the average of the inlet and outlet pressures. Here, M T is the total mass
flow rate through the channel. The conductance values obtained in the experiments by
Maurer et al [2003] and by the CNSE are also shown in the figure for comparison. As
shown in Figure 4.10, the classical equations do not predict the paradox at all and the
conductance decreases continually with increasing Knudsen number. However, the
extended equations predict the ‘Knudsen paradox’ and the results also excellently

59
match with the experimental measurements. The physical reasons for the presence of
the ‘Knudsen paradox’ are given below.

CNSE
ENSE
Experiment_Maurer et al [2003]

Figure 4.10 Conductance of the microchannel, Cˆ  M T Pin  Po  as a function of


average Knudsen number depicting the ‘Knudsen paradox’; Comparisons with
experimental measurements of Maurer et al [2003]

As the Knudsen number increases, the significance of diffusion mass transport in


comparison to convection increases dramatically as shown by equation (4.39). As
mentioned in sections 4.4.1 and 4.4.2, unlike the convective transport, the diffusion
mass transport does not cause any significant transport of momentum to the solid walls
because of the absence of normal diffusion velocity gradient at the wall and hence the
pressure drop due to the diffusive mass transport in the channel is negligible.
However, the convective velocity gradient in the normal direction to the wall is non-
zero and hence the convective flow results in momentum transport to the wall and
leads to pressure drop. It is interesting to note that at the critical Knudsen number
where the minimum occurs in the conductance profile, a balance is reached between
the convective and self-diffusive mass transports. Beyond this critical value, any
further increase in the Knudsen number results in an imbalance favouring enhanced
diffusion transport that requires no pressure drop in the channel. For a given pressure
difference across the channel, the presence of higher diffusion transport of mass, with
almost negligible loss of momentum to the wall, results in an enhanced mass flow
60
through the channel. Or in other words, one can obtain the same mass flow rate
obtained at a given Knudsen number with a lower pressure gradient at higher Knudsen
numbers because of enhanced diffusion transport. Therefore, the conductance of the
channel increases with increasing Knudsen number beyond the minimum value as
observed by Knudsen [1909]. This physically sound and rather simple explanation of
the ‘Knudsen paradox’, which was not readily available all these decades, cannot be
given on the basis of the CNSE. Further, the ENSE can indeed predict this
experimental observation without any empirical tuning parameters.

4.4.4 Comparison of mass flow rates

One of the important aims of any experiment involving micro-conduits is the accurate
prediction of the mass flow rate for a given pressure ratio and channel dimensions
since prior knowledge of the mass flow characteristics through the channel is
mandatory to determine the performance of any MEMS equipment. In Figure 4.11, the
experimentally measured mass flow rates through a straight microchannel by Maurer
et al [2003] are compared with the numerically obtained flow rates with the ENSE.
The mass flow rates predicted by the CNSE with the no-slip boundary condition are
also shown for comparison. Since the outlet pressure is constant in all the simulations,
a higher value of difference of square of the pressures represents lower Knudsen
number and vice versa.

Experiments_Maurer et al [2003]
CNSE
ENSE

1 2 Pin2  2
 Po2 , bar

Figure 4.11 Mass flow rate M T comparisons obtained with the CNSE and the ENSE
with the experimental measurements of Maurer et al [2003]; Po=0.12bar

It can be observed in Figure 4.11 that the CNSE predict the mass flow rates accurately
at higher values of squares of pressure, i.e. at lower Knudsen numbers. As discussed
61
earlier, the diffusion mass flow rate is negligible when the Knudsen number is small
and hence the classical equations are valid in this region. However, as the Knudsen
number increases or the difference of square of pressures decreases, the deviation in
the mass flow rates predicted by the CNSE and measured in the experiments
continually increases and even the error in the theoretical prediction reaches more than
one order of magnitude. On the contrary, the ENSE accurately predict the mass flow
rate through the microchannel at all the Knudsen numbers. This conclusively proves
that the additional mass flow rate observed in gas flow experiments in microchannels
is caused by the self-diffusion of mass.

4.5 Insights into the physics of microchannel flows

It is widely believed that the Navier-Stokes equations, even with the ‘slip-velocity’ at
the wall, cannot be used beyond a Knudsen number of 0.1 in predicting the
characteristics of gas flows through microchannels since the continuum assumption
employed in deriving these equations cannot hold at these high Knudsen number
regimes. It is also stated in the published literature that at conditions where Kn>0.1, it is
required to employ microscopic non-continuum computational tools such as the Lattice
Boltzmann Modelling (LBM) or the Direct Simulation Monte Carlo (DSMC)
techniques to obtain accurate description of the characteristics of gas flows through
microchannels; see Zhang et al [2005] and Gad-el-Hak [2002]. Interestingly, in
contrast, the results obtained with the continuum based ENSE agree excellently well
with the experimental measurements even for Kn>2, a moderately rarefied free-
molecular regime. In order to understand these apparent conflicting observations, it is
required to understand the definition of the Knudsen number.

As per the classical definition, see Gad-el-Hak [2002], the Knudsen number, defined
in equation (4.45), defines whether the fluid can be dealt as a continuum or not:
l
Kn  (4.45)
Lc
where l is the molecular mean free path and Lc the characteristic length of the flow.

Needless to mention, it is imperative that one needs to choose the right characteristic
length in equation (4.45) in order to obtain meaningful insights into the dynamics of
flows. Conventionally, in all internal flow cases, the hydraulic diameter for circular or
near-circular geometries and the cross-stream height for two dimensional channels are
chosen as the characteristic length for estimating dimensionless parameters such as
Reynolds number. Based on convention, the cross-stream height of the channel is also
employed as the characteristic length to define the Knudsen number in microchannel
flows, as shown in equation (4.4). Another definition, which is rather not used because
of difficulties in its determination, employs the gradient of pressure or density as the
characteristic length, as defined in equation (4.5); see Gad-el-Hak [1999]. As one can
observe, it may be difficult to estimate the local characteristic length scale employing
equation (4.5) in experimental measurements. Now it is necessary to choose from the
62
two definitions, the correct characteristic length of the gas flows through microchannel
that describes the flow characteristics properly.

1.8E-06 1.6
Pressure ratio
1.6E-06 Molecular mean free path 1.667 3.333 10 1.4
Local Knudsen number 10
Molecular mean free path, m

1.4E-06 1.667 3.333


1.2

Local Knudsen number


1.2E-06
1
1.0E-06
0.8
8.0E-07
0.6
6.0E-07
0.4
4.0E-07

2.0E-07 0.2

0.0E+00 0
0 0.002 0.004 0.006 0.008 0.01
Streamwise distance from the inlet, m
(a)
0.025 1.2E-03
Pressure ratio
Diffusion length scale 10
1.667 3.333
Local Knudsen number 1.667 10 1.0E-03
0.02 3.333
Diffusion length scale, m

Local Knudsen number


8.0E-04
0.015

6.0E-04

0.01
4.0E-04

0.005
2.0E-04

0 0.0E+00
0 0.002 0.004 0.006 0.008 0.01
Streamwise distance from the inlet, m

(b)

Figure 4.12 Comparison of Knudsen number at various pressure ratios; Po=0.12 bar
(a) Variation of mean free path and local Knudsen number Kn  l H ; (b) Variation of
1
 1 P  l P
inverse diffusion length scale Lc    and local Knudsen number Kn 
 P x1  P x1

63
In Figure 4.12(a), the molecular mean free path, l and the local Knudsen number
Kn  l H are plotted as a function of the streamwise coordinate for various pressure
ratios. In all the cases, the outlet absolute pressure was kept at a constant value of 0.12
bar. As observed in this figure, the molecular mean free path and local Knudsen
number profiles are similar and the maximum local Knudsen number value at the
outlet of the channel for all the cases was found to be 1.42, since the outlet pressure
was kept the same in all the simulations. It is interesting to note that as per the
classification of gas flows into different regimes based on Knudsen number, see Gad-
el-Hak [1999], the Knudsen number at the outlet in Figure 4.12(a) falls under the
category of transition or moderately rarefied regime. As stated earlier, the continuum
assumption must have seized to provide any meaningful results beyond a Knudsen
number of 0.1. Since one can obtain excellent results with the ENSE even beyond a
Knudsen number of unity, one can conclude that the Knudsen number definition
employed in Figure 4.12(a) does not describe the characteristics of microchannel
flows properly.

The local Knudsen number defined with the characteristic length given by equation
(4.4) is shown as a function of streamwise coordinate is shown in Figure 4.12(b).
1
 1 P 
Further, the diffusion length scale defined as Lc    is also plotted in the same
 P x1 
figure. It can be observed in this figure that the diffusion length scale is much higher
than the molecular mean free path l shown in Figure 4.12(a). Therefore, one can
understand that the self-diffusion present in microchannels influences flow
characteristics over longer distances in comparison to the length scales of mean
molecular motion. Evidently, one can observe in Figure 4.12(b) that the Knudsen
number defined by equation (4.5) is less by three or four orders in comparison to the
conventional definition, shown in Figure 4.12(a). Hence, one may conclude that the
height of the channel may be used as a characteristic length in microchannel flows
when the flow is dominated by only convection. However, when the flow
characteristics in micro-conduits are also determined by diffusion, the height of the
channel cannot be used to characterize the flow. Further, it is interesting to observe
that the diffusion does not vary in the cross-stream direction and is only the function
of the streamwise direction. This also points out that the characteristic length has to
be chosen along the streamwise direction.

Typically, the Knudsen number is a function of only the local pressure when the
characteristic length is taken to be the channel height. As seen in Figure 4.13(a), the
Knudsen number at the end of the channel is constant for all values of pressure ratio
since the outlet pressure has been fixed to be the same value in all the simulations.
However, when the Knudsen number is defined based on the characteristic length
presented in equation (4.5), it is a function of both the local value of pressure and its
gradient in the channel. For example, it can be observed in Figure 4.13(b) that the
local Knudsen number for the case with a pressure ratio of 10 is less near the inlet of
the channel in comparison to that of the cases with pressure ratios of 1.67 and 3.33.
64
However, towards the outlet of the channel, the local Knudsen number with the
pressure ratio of 10 is one order of magnitude more than that of the other two cases.

(a)

(b)
Figure 4.13 Comparison of conductance, Cˆ  m T Pin  Po  of the microchannel as a
l l P
function of local Knudsen number; (a) Kn  and (b) Kn 
H P x1

Hence, one can observe that the local value of the absolute pressure alone does not
determine the Knudsen number of the flow and the local pressure gradient is also
equally important. From the above discussions, it can be concluded that the
conventional definition of Knudsen number does not provide precise insight into the

65
characteristics of gas flows through microchannels. Further, the definition based on
the characteristic length presented in equation (4.5) suggests that the continuum
assumption can still be valid in most of the gas flow situations experienced in
microchannels. It is proposed that the gas flow characteristics through micro-conduits
need to be carefully revisited once again in the light of above discussions.

One can summarize the discussions presented in this chapter as given below. The
convective acceleration terms in the Navier-Stokes equations can be neglected in the
analysis of low Reynolds number creeping flow through the microchannels. The
ENSE predict all the characteristics of gas flows through microchannels accurately
without any empirical tuning parameters and provide physically sound explanations.
Hence, it can be taken as the proof that the additional mass flow rate observed in the
experiments is the self-diffusive mass transport. The classical definition of Knudsen
number does not characterize the gas flows through microchannels properly and the
diffusion length is the correct length scale to be used in microchannel flows.

66
Chapter 5

ANALYTICAL TREATMENTS OF GAS FLOWS


THROUGH MICRO-CONDUITS

5.1 Introduction

As shown in the previous chapter, the extended Navier-Stokes equations (ENSE) were
successfully employed to numerically solve gas flows through microchannels and
excellent agreements were obtained with the experimental measurements. These results
conclusively proved that the excess mass flow rate, observed in experiments, was
indeed caused by the self-diffusion of mass driven by the pressure gradient along the
length of the channel. Further, the discussions on the right definition of the Knudsen
number also clarified that the continuum-based ENSE are very much valid even at
‘apparently high’ Knudsen numbers defined by the mean molecular mean free path and
the height of channel. Buoyed up by these interesting and revealing results, an
analytical solution for the flow of gases through micro-conduits was attempted, as
explained in this chapter. Needless to mention, the availability of an analytical solution
to determine the flow characteristics of microchannels and capillaries will completely
eliminate the high computational efforts required in the currently employed simulation
techniques, such as the Lattice Boltzmann Method (LBM) or the Direct Simulation
Monte Carlo (DSMC) computations.

In this chapter, the analytical solution procedure employing the ENSE is explained in
detail and a similar treatment is also available in the articles co-authored by the author;
see Filimonov et al [2010] and Navaneetha Krishnan et al [2010]. The comparisons of
mass flow rates of gas flows through microchannels obtained based on the analytical
solutions are carried out with the numerical solutions described in chapter 4 and the
experimental measurements by Maurer et al [2003], Arkilic et al [1994] and Colin et al
[2004] are shown. Furthermore, a semi-analytical solution of the pressure profile along
the length of the channel is also obtained and the same is compared with corresponding
results of the numerical solutions and the experimental measurements. Furthermore, the
analytical results obtained for micro-capillary flows are also compared with the
experimental measurements by Yang and Garimella [2009]. In addition, a characteristic
pressure is defined, based on the dimensions and fluid properties for both microchannels
and micro-capillaries and it is shown to be a very important parameter in determining
the characteristics of gas flows through micro-conduits.
5.2 Analytical solution procedure

5.2.1 Order of magnitude analysis

In the analytical solution procedure, the gas flow is assumed to be two-dimensional for
microchannels and axisymmetric for capillaries in rectangular and cylindrical
coordinate systems, respectively. Furthermore, the flows are also assumed to be
steady and isothermal in nature. Since the fluid is assumed to be isothermal, the
viscosity is also considered to be constant. The schematic representation of the
convective and total velocity profiles obtained from the numerical solutions of the
CNSE and the ENSE mentioned in the previous chapter are shown in Figure 5.1. The
ENSE in the total velocity form, obtained from equations (3.46) and (3.49), can be
written for steady, isothermal gas flows, as given below:

Continuity equation:
U iT 
0 (5.1)
xi
Momentum equations:
 P   C 2 D C

xi
 
U iT U Tj   
x j xi
D D
 ij  m i U j  3  ij m k U k  (5.2)

The density is calculated based on the equation of state given by equation (2.3).
Further, equations (5.1) and (5.2) can be expanded for a two-dimensional flow
situation, employing equation (2.24), as

Continuity equation:
U1T  U 2T 
 0 (5.3)
x1 x2

x1 - momentum equation:
 4 U 1C 2 U 2C 
     m 1DU 1D 
  
 U 1T U 1T

 U 1T U 2T  
P


  3 x1 3 x 2 
x1 x 2 x1 x1  2 D C 

 3

 m 1 U 1  m 2DU 2C  
(5.4)

  U 1C U 2C D D 
      
m U 
x 2  
2 2
x 2 x1 

68
x 2 - momentum equation:
  
 U 1T U 2T

 U 2T U 2T 

P

 
 
U 1C
 
U 2C
 
m D
U D 

x 2 x1  
1 2
x1 x 2 x 2 x1 
  4 U 2C 2 U 1C 2 

x 2
     
 m 2DU 2D  m 1DU 1C  m 2DU 2C  (5.5)
 3 x 2 3 x1 3 

x2 or r

x1

(a)
x2 or r

x1

(b)
Figure 5.1 Schematic representation of velocity profiles obtained from the numerical
solutions presented in chapter 4; (a) Convective velocity from the CNSE (b) Total
velocity from the ENSE

For the case of gas flows through straight microchannels, there is no diffusion
transport of mass in the cross-stream direction since the pressure is constant in this
direction. Similarly, the convective velocity in the cross-stream direction is also zero
since the flow is assumed to be fully developed. Further, as shown in chapter 4, the
convective acceleration terms can be neglected in the momentum equations for the
case of gas flows through micro-conduits. Therefore, equations (5.3) to (5.5) can be
simplified as follows:

69

 U1T
0
 (5.6)
x1
P  4 U 1C
 D D 2 D C   U 1C 
0     
m 1 U 1  
m 1 U 1   
   
 (5.7)
x1 x1
 3 x1 3  x2  x 2 
P   U 1C    2 U 1
C
2 
0   

x 2 x1 

  
x 2  x 2  3 x1 3
 
 m 1DU 1C 
(5.8)

To further simplify equations (5.6) – (5.8), an order of magnitude analysis was


employed. The characteristic velocity and length scales for this analysis had to be
carefully chosen. The convective velocity can be scaled with the average velocity at
the exit of the channel U and the diffusion velocity is scaled as ~   L based on the
expression for the self-diffusion velocity given by equation (3.43). Further, the
characteristic length for the streamwise and cross-stream directions are the length L
and height H of the channel, respectively where H<<L. Furthermore, one can scale the
pressure with the dynamic pressure U 2 . Subsequently, one can perform the order of
magnitude estimates for the various terms in equation (5.7) as given below.
P U 2
~ (5.9a)
x1 L
  4 U 1C  U
  ~ 2 (5.9b)
x1  3 x1 
 L
 2
x1
 
 m 1DU 1D ~ 3 (5.9c)
L
  2 D C U
  m 1 U 1  ~  2 (5.9d)
x1  3  L
  U 1C  U
   ~ 2 (5.9e)
x2  x 2  H
 

Neglecting the smaller terms in equations (5.9), equation (5.7) can be rewritten as:
P   U 1C 
0    (5.10)
x1 x 2  x 
 2 

Similarly, the order of estimate analysis of the various terms in equation (5.8) is given
below:
P U 2
 ~ (5.11a)
x 2 H

70
  U 1C  U
  ~
x1  x 2  HL (5.11b)
 

  2 U 1C  U
  ~
x 2  3 x  HL (5.11c)
 1 

  2 D C  U
x 2
 
 3 m 1 U 1  ~  HL (5.11d)

Neglecting the smaller terms in equation (5.11), equation (5.8) can be reduced to the
following form:
P
0 (5.12)
x 2

As observed from the numerical solutions given in chapter 4, the pressure was found
to be only a function of the streamwise coordinate x1 in micro-conduits and it is also
evident from equation (5.12). It is well known that the diffusion velocity, defined by
equation (3.43), is a function of only the streamwise coordinate x1 and hence one can
write the following expression:
U 1D
0
x 2 (5.13)

Employing equation (5.13), equation (5.7) can be written as:


P   U 1C    U 1D 
     0 (5.14)
x1 x 2  x  x  x 
 2  2  2 

For isothermal gas flows through microchannels, the dynamic viscosity  is a constant
and hence equation (5.14) is reduced to:
P   U 1T 
   0 (5.15)
x1 x2  x 
 2 

The final set of simplified governing equations for gas flows through microchannels is
given below:
U 1T 
0 (5.6)
x1
P   U 1T 
   0 (5.15)
x1 x 2  x 
 2 

71
P
0 (5.12)
x2

Similarly, the following governing equations for gas flows through capillaries can also
be derived:
U1T 
0 (5.6)
x1
P 1   U 1T 
  r 0 (5.16)
x1 r r  r 

P
0 (5.17)
r

5.2.2 Analytical solutions of microchannel flows

One can observe that since the total velocity is a function of only the streamwise
coordinate x1 , the continuity equation is always satisfied. Hence, to obtain an
analytical solution for the fully developed gas flows through microchannels, one can
integrate the momentum equation, shown in equation (5.15), subjected to the
following boundary condition:
 dP
U 1T x1    at x 2   h (5.18)
P dx1
where h is the half-height of the microchannel, i.e. h  H 2 . On integrating equation
(5.15) twice with respect to the cross-stream coordinate x 2 , the following expression
for the total velocity is obtained:
1 dP 2
U 1T  x1 , x 2   x2  Ax2  B (5.19)
2 dx1

Employing the boundary conditions given in equation (5.18), the integration constants
A and B can be determined and hence, equation (5.19) is written as:
1 dP 2   1 dP 
U 1T  x1 , x 2   
2  dx1
 
h  x 22   
  P dx1 
(5.20)
      
Term - I Term - II

Equation (5.20) can be employed to calculate the total velocity at any location of the
microchannel, where Term–I represents the parabolic profile of the convective
velocity through the microchannel and Term–II is the block profile due to the self-
diffusion mass transport. Further, the mass flow rate profile can be written as:
 dP 2  dP
U 1T x 2   
2 dx1
h  x22  
P dx1
(5.21)

72
It should be noted that the second term on the right-hand side of equation (5.21) is the
additional self-diffusive mass flow rate. Furthermore, it can also be observed from
this equation that the diffusion mass flow rate will assume significance only in the
case of low pressures and high pressure gradients. Further, it can be observed in
equations (5.20) and (5.21) that the total mass flow rate profile is only a function of
the cross-stream coordinate whereas the total velocity profile is a function of both the
cross-stream and streamwise coordinates.

Based on equation (5.21), one can also derive the relationship between the total mass
flow rate m T , the diffusive mass flow rate m D . The expression for the total mass flow
rate through the cross-section of the channel can be obtained by integrating equation
(5.21) along both the streamwise and cross-stream directions. The integration of
equation (5.21) in the cross-stream direction leads to:
h h
  dP 2  dP 
M 1T  w  U 1T dx2   w   h  x 22   dx 2 (5.22)
h  h  2  dx1 P dx1 
where w is the width of the microchannel. On integrating and employing the boundary
conditions given by equation (5.18), equation (5.22) is reduced to:
 h 2    dP
M 1T  2hw   (5.23)
 3 P dx
 1

By employing the equation of state given by equation (2.3), equation (5.23) can be
rewritten as:
 h2 P   dP
M 1T  2hw   (5.24)
 3 T P dx
 1
At every streamwise location x1 in the microchannel where the local pressure P and its
gradient dP dx1  are known, the total mass flow rate M 1T can be computed
employing equation (5.24). Furthermore, from the second term of this equation, one
can see that the diffusion component of the velocity profile will only become
significant at low pressures and/or high temperatures. Hence, to achieve
”microchannel effects” at room temperatures, the pressure at the exit of the
microchannel has to be chosen very low and only then the influence of self-diffusion
will be present in the flow. Based on this expression, it is possible to explain the
strong deviation from the classical theory obtained by Maurer et al [2003] and nearly
no deviation in the work of Colin et al [2004].

Further, integrating equation (5.24) in the streamwise direction, from the inlet x1  0
to the outlet  x1  L , the complete analytical expression for the total mass flow rate
through the microchannel can be obtained as:
L
T
L 
 h2 P   dP 
M
 1 1 dx   2 hw   3T  P  dx dx1 (5.25)
0 0   1

73
The limits of the integration can be changed to pressure values at the boundary
conditions as:
Po
L
T  h2P 
 M 1 dx1  2hw   3T  P dP (5.26)
0 Pin  

On simplifying, the following expression can be obtained from equation (5.26):


2
h 3 w Po2  Pin 
T
 2  h w  P 
M  
 
  1  ln  in  (5.27)
3 L  T  Po   L Po 
    
Term - II

Term - I

In equation (5.27), the subscript ‘1’ has been dropped in the total mass flow rate since
it remains constant along the length of the channel. It is evident that the total mass
flow rate M T through the channel can be expressed as a function of the pressure ratio
PR  Pin Po  . Furthermore, one can further generate more meaningful results from
the expression for the mass flow rate, given by equation (5.24), by rearranging the
equation, as:
 dP  h 2 P 2 
M T  2hw  2  1 (5.28)
P dx1  3 T 

dP
Since the local pressure P and its gradient are only functions of the streamwise
dx1
direction x1 , the diffusive mass flow rate through the channel M 1D can be expressed as:
 dP  h  dP
U 1D   w  dx2  2hw (5.29)
P dx1 h P dx1

Substituting equation (5.29) in equation (5.28), the expression for the ratio of mass
flow rates can be obtained as:
MT  h2 P2 
  2  1 (5.30)
M 1D  3  T 

On rewriting, equation (5.30) becomes:


M 1D 1
T
 2 2
(5.31)
M  h P 
 2   1
 3 T 
By observing equation (5.31), it is possible to define a characteristic pressure PC as:

PC  3T (5.32)
h

By substituting the definition of the characteristic pressure given in equation (5.32),


equation (5.31) is rewritten as:
74
M 1D 1 1
T
 2
 2 (5.33)
M P PC   1 P  1
A universal relationship between the ratio of self-diffusion and total mass flow rates
and the normalized local pressure, presented in equation (5.33), was found to be an
important tool in characterizing gas flows through microchannels, as discussed later in
this chapter.

5.2.3 Analytical solutions of capillary flows

To obtain an analytical solution for fully developed gas flows through capillaries, one
can solve equation (5.16) subjected to the following boundary conditions. The first
boundary condition is that the total velocity U 1T is equal to the diffusive velocity U1D
at the solid wall, since the convective velocity becomes zero, as explained in Figure
4.6:
 dP
U 1T  x1    at r  R (5.34)
P dx1
where R is the radius of the considered micro-capillary. Further, it can be observed
that the total velocity U 1T is a continuously differentiable function with respect to r.
Therefore, it is possible to write the following expression as the second boundary
condition:
U 1T x1 , r 
 0 at r  0 (5.35)
r

On integrating equation (5.16), one can obtain the expression for the total velocity as:
1 dP 2
U 1T  x1 , r   r  Ar  B (5.36)
4 dx1

Further, by substituting the boundary conditions given by equations (5.34) and (5.35),
the constants A and B in equation (5.36) can be evaluated and the expression for the
total velocity U 1T , given by equation (5.36), can be rewritten as:
1 dP 2  dP
U 1T x1 , r   
4  dx1
R  r 2 
P dx1
(5.37)

Further, the total mass flow rate profile can be expressed as:
 dP 2  dP
U 1T r   
4  dx1
R  r 2 
P dx1
(5.38)

The analytical expression for the total mass flow rate through the capillary can be
obtained by integrating equation (5.38) in the radial and streamwise directions. The
integration in the radial direction can be expressed as:

75
R R
  dP  dP 
M 1T  2  U  rdr  2   4 dx R 
T 2
1  r2 
 rdr (5.39)
0 0  1 P dx1 
On integrating, equation (5.39) can be simplified as:
   dP
M 1T  R 2  R 2   (5.40)
 8 P  dx1

Employing the equation of state given by equation (2.3), equation (5.40) can be
reduced to:
 P   dP
M 1T  R 2  R2   (5.41)
 8T P  dx1

Integrating equation (5.41) in the streamwise direction from the inlet x1  0 to the
outlet  x1  L, one can obtain the complete analytical expression for the mass flow rate
of gases through capillaries, as given below:
L L
T 2  P 2   dP 
 M 1 dx1  R   8T R  P  dx  dx1 (5.42)
0 0   1

Further, equation (5.42) can be rewritten as:


L Po
T 2  P 2 
M
 1 1 dx  R   8T R  P  dP (5.43)
0 Pin  

On integrating, equation (5.43) becomes:


MT R2 P 
2
L
16T
 
Pin2  Po2   ln in  (5.44)
R  Po 
from which the final equation for the total mass flow rate of gases through capillaries
can be written as:
R 4 Po2  Pin2  R 2  Pin 
MT   2  1  ln  (5.45)
6TL  Po 
 L  Po 
 

Term - I Term - II

Equation (5.45) can be employed to calculate the total mass flow rate through
capillaries. By comparing equations (5.45) and (4.34), one can easily observe that the
first term is the result of the convective mass transport and can also be obtained with
the help of the CNSE. The second term is due to the diffusion mass transport obtained
with the ENSE. Similar to the analysis shown in section 5.2.1 for the case of
microchannel flows, one can obtain the relationship between the total and diffusive
mass flow rates as:
2 2
T 2  dP  R P 
M  R  2  1 (5.46)
P dx1  8 T 

76
The diffusive mass flow rate through the capillary can be written as:
 dP
M 1D  R 2 (5.47)
P dx1

Employing equation (4.47), equation (5.46) can be rewritten as:


M 1D 1
T
 2 2 (5.48)
M R P
1
8 2 T

By observing equation (5.48), it is possible to define a characteristic pressure PC for


capillary flows as:

PC  8T (5.49)
R

Substituting the definition of the characteristic pressure given in equation (5.49),


equation (5.48) can be expressed as:
M 1D 1 1
T
 2
 2 (5.50)
M P PC   1 P  1

Equation (5.50) is the universal relationship between the ratio of local diffusive mass
flow rate to the total mass flow rate and the local normalized pressure in gas flows
through capillaries. Further, it can also be used to determine whether the diffusive
mass flow rate needs to be considered in the analysis for the given geometry and
boundary conditions.

5.2.4 Semi-analytical solutions of pressure in microchannels and capillaries

As mentioned in the previous subsection, the mass flow rate through microchannels
and capillaries can be determined if the geometry and the static pressures at the inlet
and outlet locations are given. The pressures at the boundary locations such as inlet
and outlet of the conduit can be measured easily. However, there are no analytical
solutions available for determining the pressure profile along the length of
microchannels or capillaries. Needless to mention, if this solution is made available, it
can help in determining the flow characteristics through micro-conduits completely. It
is for this reason that, in this section, the governing equation for the pressure profile
along the flow direction is derived to yield the required pressure and pressure gradient
information of the flow.

The mass flow rate M T through a microchannel is given by equation (5.24). Due to
conservation of mass, M T is constant along the streamwise direction x1 . Therefore,
one can write

77
d
dx1
 
MT 0 (5.51)

Equation (5.51) is nothing but the continuity equation, presented in equation (5.6).
Substituting the expression for the total mass flow rate M T through microchannels
from equation (5.24), equation (5.51) can be modified to result in an equation for the
pressure distribution along the microchannel as:
d   h2P   dP 
  2 hw   0 (5.52)
dx1   3  T P dx
 1

Equation (5.52) can be expanded to yield the following equation after differentiation
with respect to x1 , as given below:
 dP  h 2 dP  dP  d 2 P  h 2 P  
 
 3T dx  2

 dx 2  3T P    0
   (5.53)
dx
 1 1 P dx1  1  

Equation (5.53) is the governing equation for the pressure profile in the fluid along the
flow direction of microchannels with the boundary conditions Pin and Po at the inlet
and outlet, respectively, specified according to the experimental conditions. Similarly,
the mass flow rate profile for the case of capillary flows, shown in equation (5.41), can
be employed in equation (5.51) to yield the equation for the pressure distribution as:
 dP  R 2 dP  dP  d 2 P  R 2 P  
   2   
2 
    0 (5.54)
 dx1  8T dx1 P dx1  dx1  8T P  

In order to obtain the semi-analytical solutions for the boundary value problem
represented by equations (5.53) and (5.54), these equations were solved iteratively
using a 4 th order Runge-Kutta method to obtain the pressure and pressure gradient
profiles along the micro-conduit. These results were employed to calculate the total
mass flow rates using equations (5.28) and (5.46) for different inlet and outlet pressure
values. Further, the results thus obtained are compared later with the experimental
measurements of Maurer et al [2003] and the numerical results shown earlier in
chapter 4.

5.3 Results and Discussions of gas flows through microchannels

5.3.1 Regimes of flows in microchannels

The profile of the mass flow ratio given by equation (5.33) is shown in Figure 5.2 as a
function of the pressure ratio P  P PC  . From equation (5.33) and Figure 5.2, one
can see that when P  PC , the diffusive mass flow rate M 1D is equal to 50% of the total
mass flow rate M T , i.e. the convective and the diffusive mass flow rates are of the
same magnitude. As the pressure decreases along the flow direction of microchannels,
78
the convective mass flux dominates when P  PC and subsequently, the diffusive mass
flux becomes the major component of the total mass flux when P  PC . This
observation can further be elaborated as given below.

Table 5.1 Summary of experimental conditions employed by Maurer et al [2003],


Arkilic et al [1997] and Colin et al [2004]

Maurer Arkilic Colin et al [2004]


Experimental parameters et al et al
[2003] [1997] Case a Case b Case c Case d
Reported channel height H
μm 1.14 1.33 1.88 1.88 1.88 1.88

Calculated channel height


1.14 1.41 1.94 2.06 1.94 2.06
H’ μm
Width w μm 200 52.25 21.2 21.2 21.2 21.2
Length L μm 10000 7500 5000 5000 5000 5000
Gas used Helium Helium Nitrogen Nitrogen Helium Helium
Gas constant [J/Kg.K] 2077 2077 297 297 2077 2077
Temperature [K] 293 314 294.2 294.2 294.2 294.2
Viscosity of the gas
1.99 2.066 1.9 1.9 1.75 1.75
[10-5 Pa s]
Characteristic pressure
47.7 43.5 10.03 9.45 24.43 23.0
[kPa]
160-
Inlet pressure range [kPa] 26-507 400 400 400 400
420
Outlet pressure [kPa] 12-101 100 200 65 190 102.6
Outlet Knudsen number 1.509-
Kno  0.157 0.017 0.048 0.056 0.098
0.179

In Figure 5.2, the pressure ranges of the experimental data of Maurer et al [2003],
Arkilic et al [1994] and Colin et al [2004] are also indicated and the experimental
conditions employed in these publications are tabulated in Table 5.1. As mentioned
earlier in chapter 4, Maurer et al [2003] obtained significant deviations in their
experimental measurements in comparison to the predictions based on the CNSE and
the gap between the experimental and theoretical results were bridged by introducing
the slip-velocity, as described in section 4.1. As repeatedly stated in chapters 4 and 5,
the additional mass flow rate observed in the gas flow experiments through
microchannels is nothing but the self-diffusion transport of mass due to the presence
79
of strong pressure gradients. One can make the following observations based on the
dimensionless pressure ranges, obtained based on the characteristic pressure PC ,
indicated in Figure 5.2. In the pressure range of Arkilic et al [1994], the diffusion
effects are present but they are quite small while in the case of Colin et al [2004],
there are no diffusion effects at all and the results obtained with the CNSE and the
ENSE must be identical. The discussions given above suggest that the characteristic
pressure can be an excellent tool in characterizing gas flows through microchannels.

5.3.2 Discussions on pressure profiles

Before discussing the various results obtained with the analytical solutions for
microchannel flows, one can make certain interesting interpretations about the
characteristic pressure shown in equation (5.32) and the equation of the pressure
profile shown in equation (5.53). The governing equation of the pressure profile along
the microchannel, represented by equation (5.53), can be rewritten by introducing the
characteristic pressure PC , given by equation (5.32), as given below.
 dP   dP  dP  d 2 P   P  
 
 2  

 2    0
2 
(5.55)
2
 dx1  PC dx1 P dx1  dx1  PC P 

One can observe that the equation of the pressure profile for the case of micro-
capillary flows, given by equation (5.54), is also reduced to equation (5.55). Since the
dynamic viscosity is a constant in isothermal gas flows through micro-conduits,
equation (5.55) can be rewritten as:
 
 dP  2  1 1  d P  P 1 
2
   2  2   2  2    0 (5.56)
dx
1   PC
 
P  dx1  PC P 
 
 Term - I Term - II 

One can observe from equation (5.56) that when P  PC , the first term of the equation
d 2P d 2P
becomes zero and one obtains  0 since the coefficient of in the second
dx12 dx12
term is always positive. One can further elaborate on the above mentioned
observation and state that the streamwise distribution of pressure in micro-conduits
has a point of inflection when the local pressure is equal to the characteristic pressure.
dP
Therefore, one can further note that the pressure gradient reaches a minimum
dx1
when the local pressure is equal to the characteristic pressure PC . Since the pressure
gradient is always negative in gas flows through straight micro-conduits, i.e. the
pressure decreases along the length of the channel, one can make the following
observations on the behaviour of the pressure profile based on equation (5.56). For
d 2P
P  PC , one has  0 and therefore the pressure profile has to be convex in nature
dx12
80
d 2P
in this region. When P  PC , one can obtain  0 and hence, the pressure profile
dx12
becomes concave in this region.

1.0

I II III
0.8
Mass flow ratio, M1D/MT

0.6

Diffusive flux dominates Convective flux dominates


0.4

0.2

0
10-2 10-1 100 101 102

P  P / PC

Figure 5.2 Ratio of the diffusion to total mass transports through microchannels
obtained by employing equation (5.33); Pressure ranges of the experimental
measurements are shown as I - Maurer et al [2003], II - Arkilic et al [1994] and
III - Colin et al [2004]

Therefore, for a given set of inlet and outlet pressures in a microchannel, Pin and Po ,
respectively, one can make the following observation. If both Pin and Po are less than the
characteristic pressure PC , i.e. Pin  PC  Po , a complete convex pressure profile will be
d 2P
obtained, i.e.  0 , throughout the channel. Since the inlet and outlet pressure
dx12
values are less than the characteristic pressure, one will obtain strong diffusive
influences in the entire channel flow. On the other hand, when the inlet and outlet
pressures are more than the characteristic pressure, i.e. Pin  Po  PC , one would obtain
d 2P
a concave pressure profile, i.e.  0 and the influence of diffusion on the total flow
dx12
rate will be negligible or small, depending on the chosen inlet and outlet pressures.

81
Analytical solutions, CNSE

Numerical solutions, ENSE Analytical solutions, CNSE

Semi-analytical solutions, ENSE Numerical solutions, ENSE

Semi-analytical solutions, ENSE

x1 / L x1 / L
(a) (b)

Analytical solutions, CNSE

Numerical solutions, ENSE

Semi-analytical solutions, ENSE

Analytical solutions, CNSE

Numerical solutions, ENSE

Semi-analytical solutions, ENSE

x1 / L x1 / L
(c) (d)

Figure 5.3 Comparison of pressure profiles obtained with the CNSE and the ENSE
along the length of the microchannel; PC  41483.5 Pa ; Po  0.29PC ; PR  Pin Po
Pin PC  valuesgiven in brackets (a) 58.3 (16.9); (b)10 (2.9); (c) 3.33 (0.97) and
(d) 1.67 (0.48)

Further, let us consider the case where the characteristic pressure lies between the inlet
and outlet pressures, i.e. Pin  PC  Po . Since the characteristic pressure PC is reached
somewhere inside the microchannel, there is a point of inflection in the pressure
profile along the length of the channel where the local pressure is equal to the
characteristic pressure. At this point, the pressure profile moves towards the convex
profile from the concave profile. In this case, the convection will dominate the
diffusion transport until the location where the local pressure is equal to the
characteristic pressure and the diffusion will play a dominant role in determining the
total mass flow rate in the rest of the length of the microchannel.

82
One can refer to the pressure characteristics obtained by the numerical solutions based
on the ENSE, shown in Figure 4.9 and observe the supporting evidence for the
arguments presented above. As it can be observed from Figure 4.9, the pressure
profiles change from concave to convex with decreasing pressure ratio and in Figure
4.9(b), one can also see the point of inflection in the pressure profile. This is an
encouraging observation in order to make a quantitative comparison between the semi-
analytical and numerical solutions obtained with the ENSE.

In Figure 5.3, the semi-analytical solutions for the pressure profile along the length of
the microchannel obtained with the CNSE and the ENSE are shown for different inlet
to outlet pressure ratios. Further, the numerical solutions described in Figure 4.9 are
also plotted in the same figure for comparison. The outlet pressure was kept to be
constant in all the simulations. The local dimensionless pressure is defined as
P  P  Po  Pin  Po  so that one can get 0  P  1 for ease of comparison at different
pressure ratios across the microchannel. Similar to Figure 4.9, one can also observe in
Figure 5.3(a) that when the pressure ratio is high, there is no difference in the pressure
profiles between the semi-analytical solutions obtained with the CNSE and the ENSE.
The numerical results obtained with the ENSE also agree very well with the analytical
results. Since the convective mass transport dominates the flow characteristics in most
parts of the channel, one can neglect self-diffusive transport of mass in this case and
this is evident by the concave shape of the pressure profile.

However, one can observe in Figure 5.3(b) that though the convection still dominates
the flow characteristics near the inlet of the channel, the fraction of channel length
dominated by the self-diffusion near the outlet of the channel increases as the pressure
ratio is reduced. Hence, the pressure profile transforms from a concave shape near
the inlet to a convex shape towards the outlet at the point where the local pressure is
equal to the characteristic pressure of the channel, as shown in Figure 5.3(b).
Subsequently, when the inlet pressure is also less than the characteristic pressure, as
shown in Figures (5.3c) and (5.3d), the pressure profile remains convex throughout the
length of the channel and the diffusion mass transport is significant everywhere. Here,
the nature of the pressure profiles obtained with the CNSE and the ENSE are
completely different.

One can also observe in Figure 5.3 that there is an excellent agreement between the
semi-analytical results obtained in this chapter and the numerical results presented in
chapter 4. Hence, one can claim that the order of magnitude considerations carried out
in section 5.2.1 resulted in physically correct simplifications of the ENSE for gas
flows through micro-conduits. Further, one can carry out the following
nondimensional analysis to bring out more insights into the governing equation of
pressure, equation (5.56), in microchannels. To non-dimensionalize the governing
equation, the following two dimensionless quantities are employed.
P
Dimensionless pressure: P  (5.57a)
PC

83
x1
Dimensionless x1 coordinate: x*  (5.57b)
L

Employing the dimensionless quantities mentioned in equation (5.57), equation (5.58)


can be rewritten as given below.
2
 dP   1  d 2 P  1 
 *  1  2    P    0 (5.58)
 dx   P  dx *2  P 

One can observe that the length L of the microchannel drops out from the
dimensionless governing equation for the pressure distribution shown in equation
(5.58). This implies that the nondimensional pressure distribution as a function of the
dimensionless streamwise coordinate is identical for microchannels and capillaries of
all lengths if the same inlet and outlet pressures are applied. This result further
reconfirms that the characteristic pressure-based scaling introduced in the analytical
solution procedure is correct and valid.

5.3.3 Discussions on pressure gradient profiles

As discussed above, the choice of the inlet and outlet pressures with respect to the
characteristic pressure of the microchannel determines the position of the inflection
point in the pressure distribution where the ‘self-diffusion effects’ dominate the flow
characteristics. Further, one could also argue by observing equation (5.55) that the
nature of the pressure gradient profile also changes before and after the inflection
point. In order to elaborate this point further, the profiles of pressure and pressure
gradients are discussed together in this section. The pressure gradients are normalized
as:
 dP   
 dP   dx    dP dx 
 1  1  min
   (5.59)
 dx1  norm  dP 
   dP 

 dx1  max  dx1  min

In equation (5.59), ‘norm’, ‘min’ and ‘max’ represent the normalized, minimum and
maximum values of the pressure gradient, respectively, along the length of the
channel. In Figure 5.4, the dimensionless pressure and pressure gradient profiles
along the length of the microchannel are plotted at the constant inlet to outlet pressure
ratio obtained with different inlet and outlet pressure combinations. As discussed
earlier, when the outlet pressure is higher than the characteristic pressure, the
dimensionless pressure profile is concave in nature. As the outlet pressure becomes
less than the characteristic pressure, the pressure profile becomes convex. Further,
one can also observe the point of inflection in the pressure profiles where the local
pressure is equal to the characteristic pressure.

84
Po = 0.25PC
Po = 0.35PC
Po = 0.5PC
Po = PC

P = 0.62 Po = 2PC

P = 0.33

Po = 0.25PC
Po = 0.35PC
Po = 0.5PC
Po = PC
Po = 2PC

Dimensionless length along the channel, x1 / L

Figure 5.4 Profiles of pressure and pressure gradient profiles along a microchannel for
different outlet pressures; PR  4

Based on the observation of the governing equation of pressure profile, equation


d 2P
(5.55), it was mentioned that becomes zero when the local pressure is equal to
dx12
85
the characteristic pressure and the pressure gradient will have the minimum at this
point. As expected, there was no point of inflection in the pressure gradient profiles
for the case of Po  2 PC since the local pressure was more than the characteristic
pressure at all the locations along the length of the channel in this case. However, one
can observe the point of inflection in the pressure gradient profiles in Figure 5.4 when
the outlet pressure becomes Po  0.25 PC , 0.35 PC , 0.5PC and PC . As stated above, the
diffusive mass flux starts to dominate the convective mass transport when the local
pressure in the fluid reaches the characteristic pressure and hence, the behaviour of the
pressure profile changes at this point. When Po  PC , the point of inflection occurs
exactly at the outlet whereas when Po  0.25PC , it occurs at the inlet of the
microchannel. Interestingly, as the pressure ratio increases, the point of inflection in
the pressure gradient profiles moves from the inlet towards the outlet of the channel.

Based on the above mentioned discussions, the semi-analytical solutions for the
pressure and pressure gradient profiles along the length of the microchannel are very
useful in determining the local pressure values. It is stressed here that a solution for
the pressure profile along the channel is provided for the first time and no such
solution is available in the literature.

5.3.4 Comparison of total mass flow rates

With the help of the analytical solutions given in equation (5.27), one can calculate the
total mass flow rate through a microchannel. The experimental measurements of
Maurer et al [2003] with helium as the working fluid were employed to compare with
the analytical solutions. The experimental conditions employed by Maurer et al
[2003] are tabulated in Table 4.2 and also in Table 5.2. The total mass flow rates
obtained by the analytical and numerical solutions of the ENSE are compared with the
experimental measurements in Figure 5.5. The analytical solutions obtained
employing the CNSE are also shown in the same figure.

It is evident from Figure 5.5 that the solutions obtained with the ENSE agree
remarkably well with the experimental measurements and there is no difference
between the analytical and numerical solutions obtained with the ENSE. Further, the
solutions obtained with the CNSE deviate significantly from the experimental
measurements and the solutions of the ENSE in regimes where the self-diffusion mass
transport plays a significant role in determining the flow characteristics. The region
where the convection mass transport is dominant, the differences between the two
solutions are negligible and the agreement with the experimental measurements is also
good. The above comparison clearly indicates that the simplifying assumptions done
to arrive at the analytical solutions are correct and one can employ the analytical
solution provided in equation (5.27) to accurately calculate the total mass flow rate of
isothermal gas flows through microchannels.

86
Further, the total mass flow rates can be calculated for the experimental conditions
employed by Arkilic et al [1997] and Colin et al [2004] with the help of equation
(5.27). The experimental conditions employed in these publications are summarized
in Table 5.1. The analytical results obtained for the experimental conditions employed
by Arkilic et al [1997] and Colin et al [2004] are compared with the experimental
measurements as a function of the pressure ratio in Figures 5.6 and 5.7, respectively.

Numerical solutions,
Classical NS equationsCNSE
Analytical
Analytical solutions, ENSE
solutions of extended NS equations
Numerical solutions
Numerical solutions,ofENSE
extended NS equations
Total mass flow rate, Kg/s

Experimental results,
Experiments, MaurerMaurer et al. (2003)
et al [2003]

  
0.5 Pin2  Po2 Pa 2

Figure 5.5 Comparison of the total mass flow rate M T obtained by different methods
(a) Solutions based on the CNSE (b) Analytical solutions of the ENSE (c) Numerical
results of the ENSE, as shown in Figure 4.11 and (d) Experimental measurements by
Maurer et al [2003]

In Figure 5.6, the results obtained from the ENSE are shown by solid lines and those
given by the CNSE are shown by dashed lines. It can be observed in Figure 5.6 that
there are significant deviations between the theoretical results (computed with both the
CNSE and ENSE) obtained with the reported microchannel heights in the literature, as
shown in Table 5.1, and the experimental measurements. One can observe that there
is insignificant difference between the CNSE and the ENSE obtained with the reported
channel height in Figure 5.6. However, there are significant deviations between the
experimental measurements reported by Colin et al (2004) and the results obtained
with the CNSE and the ENSE, as shown in Figure 5.6.

One may claim that this deviation can be the result of the inaccuracy of the solutions
of the ENSE. However, it is to be noted that the experiments conducted by Arkilic et
87
al [1997] and Colin et al [2004] were carried out in the convection dominated regime,
i.e. Po PC   2.3 in the case of Arkilic et al [1997] and 6.8 - 20 in the case of Colin et
al [2004]. It is well known that the diffusion effects are negligible in these pressure
ranges, as shown in Figure 5.2 and Table 5.1. Hence, there should be negligible
difference between the solutions obtained by the CNSE and the ENSE and the same
can be observed in Figures 5.6 and 5.7. Hence, the deviations between the
numerically computed and experimental results cannot be attributed to the ENSE and
is most likely due to an error in measurement during the experimentation.

1.4 Arkilic et al [1997]


ENSE, reported channel height
CNSE, reported channel height
1.2
ENSE, adapted channel height
CNSE, adapted channel height
1.0
M T [10-11 Kg/s]

0.8

0.6

0.4

0.2

0.0
1 2 3 4 5
PR= Pin/Po

Figure 5.6 Comparison of the total mass flow rate M T obtained with the CNSE and the
ENSE and experimental measurements of Arkilic et al [1997] as a function of the
pressure ratio PR ; Reported channel height H = 1.33 μmand adopted channel height
H’= 1.41 μm

Interestingly, Colin et al [2004] and Arkilic et al [1997] showed that their second
order slip model had an excellent agreement with the experiments. As explained
earlier in section 4.1, the ‘slip-model’ is based on empirical correlations and relies
heavily on the accuracy of the experimental measurements from which the value of 
is deduced. Hence, one can always match the experimental measurements by adjusting
the value of  and hence in the ‘slip-flow’ models, it is possible to obtain very good
agreement with the experimental measurements.

Interestingly, Colin et al [2004] reported an uncertainty of 0.1 μm in the measured


channel height of 1.88 μm . Such large uncertainties can cause larger deviations
between the experimental data and the analytical solutions than the small errors in the
deduced ‘Maxwellian-slip’. Hence, to verify if the deviations of the theoretical
88
solutions from the experiments reported by Colin et al [2004] and Arkilic et al [1997]
are due to such uncertainties in the channel thickness, the following procedure was
adopted. The channel height was back-calculated by employing the solutions based on
the ENSE using the mass flow rate at a single experimental point each from the
experiments conducted by Arkilic et al [1997] and Colin et al [2004]. Using the
calculated channel height, the mass flow rates were estimated for all other
experimental values using the CNSE and the ENSE to observe whether all other
experimental data match with the theoretical predictions.

Experimental data Experimental data


ENSE, adapted height
Total mass flow rate MT, Kg/s

ENSE, adapted height

Total mass flow rate MT, Kg/s


CNSE, adapted height CNSE, adapted height
ENSE, reported height ENSE, reported height
CNSE, reported height CNSE, reported height

PR=Pin /Po PR=Pin /Po


(a) (b)
Total mass flow rate MT, Kg/s

Total mass flow rate MT, Kg/s

Experimental data
ENSE, adapted height
Experimental data
CNSE, adapted height ENSE, adapted height
ENSE, reported height CNSE, adapted height
CNSE, reported height ENSE, reported height
CNSE, reported height

PR=Pin/Po PR=Pin/Po
(c) (d)

Figure 5.7 Comparison of the total mass flow rate M T obtained with the CNSE and the
ENSE and experimental measurements of Colin et al [1997] as a function of the
pressure ratio PR ; Details of cases (a) to (d) are summarized in Table 5.1 along with the
reported and calculated values of channel heights

89
In Figures 5.6 and 5.7, the solutions obtained using the calculated channel height
values given in Table 5.1 are also shown employing the CNSE and the ENSE for
comparison with the experimental measurements reported by Arkilic et al [1997] and
Colin et al [2004], respectively. The new channel heights were calculated using the
experimental data at the lowest pressure ratio in both cases. It can be observed in
Figures 5.6 and 5.7 that all other experimental data points lay on the curves plotted
using the calculated thicknesses, proving that the error in determining the channel
thickness could be one of the most important factors which had caused the discrepancy
between the experiments and the analytical solutions. Thus, it is important to note that
in the ‘slip-flow’ theory, it was attempted to include even the experimental errors into
the momentum accommodation coefficient in order to match the experimental
measurements with the theoretical predictions. However, the chosen pressure range in
these experiments does not cause any significant diffusion transport and hence, there is
no ‘slip-velocity’ at the wall.

5.3.5 Discussions on the total velocity profiles

In Figure 5.8, the local mass flow rate U 1T x 2  , calculated with equation (5.21) and
the total velocity U1T  x1, x2 , calculated with equation (5.20) are shown at different axial
positions in the microchannel. The inlet and outlet pressures were taken to be
Pin  4.5PC and Po  0.2 PC . As one can observe from the extended continuity equation
(5.1), the mass flow rate is constant along the channel and hence, the area enclosed by
the profile of total mass flow rate U 1T , which is equal to the integral of the U 1T over
the cross-stream coordinate x 2 is constant for all the streamwise locations x1 . For
different streamwise positions along the channel, one can observe, in Figure 5.8, the
composition of the mass flow, consisting of the diffusive part (filled in by grey colour)
and the convective part which can be seen as the difference between the total and
diffusive mass transports. The diffusive part of the mass flow increases continuously
along the length of the microchannel. As one gets closer to the outlet of the channel,
the diffusive component of the mass flux increases while there is a corresponding
gradual reduction in the convective mass flux. In the expression for the total mass
flow rate shown in equation (5.21), the first term represents the convective mass flux
whereas the second term depicts the diffusive mass flux. Since the variation of the
local pressure in the given length is much more than that of the pressure gradient, the
influence of the diffusive mass transport is significant near the outlet.

In Figure 5.8(b), the local total velocity profiles are presented at different streamwise
positions in the microchannel. One can observe that the total velocity U 1T increases
along the channel at all the cross-stream locations with decreasing local pressure.
Even though the total mass flow rate is constant, the pressure and hence, the density
decreases along the length of the channel. Therefore, the total velocity increases as
one move towards the outlet of the channel. In Figure 5.8(b), one can observe that the
total and diffusive velocities continuously increase whereas the convective velocity
90
increases initially and then decreases subsequently when the influence of the diffusion
mass flow rate is significant. This can be explained as given below.

U1T

Position in the channel, x1/L, [%]

(a)

U1T

Position in the channel, x1/L, [%]

(b)

Figure 5.8 Profiles of total mass flow rate and total velocity in the microchannel;
Black arrows illustrate the total quantity, grey colour represents the diffusive part and
the convective part is the difference between the total and diffusive quantities;
(a) Total mass flow rate profiles and (b) Total velocity profiles

The average velocity can be expressed based on equation (5.24) as:


 h 2 T  dP
U 1T    2  (5.60)
 3 P  dx1

91
As explained in Figure 5.4, the pressure gradient decreases initially and then increases
subsequently when the condition Pin  PC  Po is satisfied. Therefore, the first term on
the right hand side of equation (5.60), which depends only on the pressure gradient,
follows a similar trend as the pressure gradient profile, shown in Figure 5.4. However,
the second term is strongly dependent also on the local pressure and becomes
dominant in comparison to the first term as the local value of the pressure becomes
low. Hence, the second term that represents the diffusive component of the velocity
increases monotonically and the velocity profiles shown in Figure 5.8(b) are obtained.

Dimensionless length, x1 L Dimensionless length, x1 L


(a) (b)

Dimensionless length, x1 L Dimensionless length, x1 L


(c) (d)

Figure 5.9 Streamwise profiles along the length of the microchannel; Pin  4.5PC ,
Po  0.2PC and PC  41484.5 Pa ; (a) Diffusion velocity, (b) Mean convective velocity
P  Po
(c) Non-dimensional pressure, P  and (d) pressure gradient
Pin  Po

In order to visualize the above mentioned argument better, the diffusion and
convection velocities along the length of the microchannel are plotted in Figure 5.9 (a)
and (b), respectively. Further, the local pressure and pressure gradient profiles are
92
also shown in Figure 5.9(c) and (d). On comparing Figures 5.9(b) and (d), one can
easily observe that the maximum convective velocity occurs at the same streamwise
location where the minimum value is obtained in the profile of the pressure gradient.
Further, one can also notice in Figure 5.9(a) that the diffusion velocity increases
rapidly close to the outlet of the microchannel since the local Knudsen number
increases because of the reduction in local pressure. Hence, the significance of the
diffusion transport increases in this region and therefore, the convection velocity
declines after reaching the maximum value beyond this point. As discussed earlier in
Figure 5.4, the minimum in the pressure gradient profile is reached where the local
pressure is equal to the characteristic pressure of the channel and this is also evident in
Figure 5.9(d).

5.3.6 Conductance of the microchannel and the ‘Knudsen-paradox’

ENSE, Semi-analytical solution


ENSE, Numerical solution using Fluent
Experiments, Maurer et al [2003]

Dimensionless average pressure P

Figure 5.10 Conductance of microchannel as a function of P  Pavg PC to compare the


analytical and numerical solutions based on the ENSE and the experimental
measurements of Maurer et al (2003)

The ‘Knudsen-paradox’, the minimum in the conductance profile when plotted as a


function of the average Knudsen number, was found to be one of the most intriguing
problems related to gas flows through micro-conduits. A number of approaches failed
to predict this experimental observation and modern computational tools such as the
Lattice Boltzmann Method (LBM) or the Direct Simulation Monte Carlo (DSMC)
93
simulations were employed to obtain the ‘Knudsen-paradox’. As explained in section
4.4.3, the numerical solutions of the ENSE could predict the ‘Knudsen-paradox’
accurately without any empirical adhoc tuning parameters. The detailed description of
the physical reasons for the ‘Knudsen-paradox’ to occur in the micro-conduit flows is
given in section 4.4.3. In this section, the conductance profile of the microchannel
flows based on the analytical solution obtained employing the ENSE is explained
along with certain other interesting observations in relation to the ‘Knudsen-paradox’.

In Figure 5.10, the conductance of the microchannel obtained based on the analytical
solution given by equation (5.27) is plotted against the dimensionless average pressure
Pin  Po
of the channel, defined as P  . Further, the experimental measurements of
2PC
Maurer et al [2003] and the numerical solutions of the CNSE and the ENSE shown in
Figure 4.10 are also illustrated for comparison. One can easily observe in Figure 5.10
that the solutions based on the ENSE predict the ‘Knudsen-paradox’ accurately. One
can also observe in Figure 5.10 that the minimum occurs at the dimensionless average
pressure of the order of unity.

To elaborate this result further, the conductance of the microchannel is shown as a


function of the average Knudsen number at different outlet pressure values. The inlet
pressure was varied to obtain different solutions for a given outlet pressure. The
solutions obtained by employing the CNSE are also shown at the outlet pressure value
of 0.1PC for comparison. It is interesting to note in Figure 5.11 that the minimum
conductance occurs at different average Knudsen numbers when the outlet pressure is
varied. As it can be observed in Figure 5.11, the average Knudsen number at which
the conductance reaches the minimum increases with increasing outlet pressure values.

In Figure 5.12, the Knudsen number at which the minimum conductance occurs in the
microchannel is plotted as a function of the outlet pressure ratio Po PC  . One can
observe in this figure that the Knudsen number at which the minimum conductance
occurs increases with increasing outlet pressure ratio. Further, the minimum
conductance of a microchannel increases as the outlet pressure declines. This means
that, for the same pressure ratio, a reduction in outlet pressure will result in lower
power needed to drive the same amount of mass flow. Based on the above mentioned
arguments, one can claim that the general observation, mentioned in literature, see
Karniadakis et al. (2005), that the minimum conductance occurs close to the Knudsen
number of unity, is no longer acceptable. Thus, the analytical solutions based on the
ENSE provide excellent agreements with the experimental measurements and can be
employed to predict the characteristics of gas flows through microchannels.

94
Conductance of the microchannel

ENSE, P0 = 0.001PC
ENSE, P0 = 0.01PC
ENSE, P0 = 0.1PC
CNSE

Average Knudsen number


Figure 5.11 Conductance of microchannel calculated at various outlet pressures as a
function of the average Knudsen number depicting the “Knudsen-paradox”

0.115
P 
Kn  0.4207 o 
min . cond
 PC 

Po PC 
Figure 5.12 Profile of the Knudsen number at the minimum conductance of the
microchannel for different Po PC 

95
5.4 Results and discussions of flows through capillaries
1

0.9 (a)

0.8 (b)
Mass flow ratio, M1D/MT

0.7 (c)

0.6 (d)

0.5

0.4

0.3

0.2

0.1

0
1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03
Pressure ratio, ℜ =P/PC

Figure 5.13 Ratio of the diffusion to total mass flow rates as the function of the
dimensionless pressure P  P PC ; Range of the experimental data of Yang and
Garimella (2009) are also shown for different cases

The analytical solutions for the mass flow rate of gas flows through capillaries are
given in section 5.2.2. In order to compare the analytical solutions with experimental
results, the experimental measurements provided in Figure 4 of the original article by
Yang and Garimella [2009] were chosen as the basis. The summary of the
experimental conditions employed by Yang and Garimella [2009] are provided in
Table 5.2. In order to estimate the relative influence of the diffusive mass transport in
the experimental conditions employed by Yang and Garimella [2009], the ratio of the
diffusion to total mass flow rates was calculated as a function of the dimensionless
local pressure P  P PC , as given by equation (5.50) and the same is plotted in Figure
5.13.

The inlet and outlet pressure ranges of the experimental conditions shown in Table 5.2
for all four cases are also indicated in this figure. One can observe that one can expect
some influence of diffusive mass transport only for the cases (c) and (d). As it can be
observed from Table 5.2, the outlet pressure of the capillary is much higher than the
characteristic pressure for all the cases except case (d). As described by equation
(5.50), the diffusive and convective mass flow rates will be equal when the local
pressure is equal to the characteristic pressure. Therefore, based on the experimental
96
conditions provided by Yang and Garimella [2009], one can expect strong diffusion
effects, at least in parts of the capillary, for case (d) only. Further, the inlet pressure in
all the cases is very high in comparison to the characteristic pressure and this will
increase the length of the convection dominated regime of the capillary. As one can
observe from equations (5.46) and (5.50), the diffusive mass transport will be much
less in comparison to the convective transport when both the inlet and outlet pressures
are higher than the characteristic pressure. Therefore, one would expect the solutions
obtained with the CNSE and the ENSE to be identical for cases (a) and (b).

Table 5.2 Summary of experimental conditions employed by Yang and Garimella


[2009]

Experimental Yang and Garimella [2009]


parameters Case a Case b Case c Case d
Reported radius R μm  50 25 10 10
Length L μm  95000 61000 33000 27000
Gas used Air Air Air Air
Gas constant [J/Kg.K] 287 287 287 287
Temperature [K] 293 293 293 273
Viscosity of the gas [10-5
1.82 1.82 1.82 1.82
Pa s]
Characteristic pressure
298.6 597.2 1493 1493
[Pa]
Inlet pressure range [Pa] 105165 105165 105165 103360
Outlet pressure [Pa] 2337 2337 2337 608
Outlet Knudsen number
Kno  0.0274 0.0548 0.137 0.54

In Figure 5.14, the total mass flow rates obtained with the analytical solutions of the
CNSE and the ENSE, given by the first term of equation (5.45) and the complete
equation of equation (5.45), respectively, are compared with the experimental
measurements of Yang and Garimella [2009] as a function of pressure ratio PR . The
results obtained with the ENSE are shown by solid lines whereas the results based on
the CNSE are depicted by dashed lines. In cases (a) and (b) in Figure 5.14, one can
observe that there is no influence of the diffusive mass flow rate on the total mass flow
rate, since both the inlet and outlet pressures are chosen to be much higher than the
characteristic pressure of the considered experiments shown in Table 5.2.

97
1.2E-07 1.2E-08
Classical
CNSE NS equations CNSE
Classical NS equations
1.0E-07 1.0E-08
Mass flow rate, Kg/s

Mass flow rate, Kg/s


ENSE
Extended NS equations ENSE
Extended NS equations
8.0E-08 Yang
Yang& and
Garimella [2009](2009)
Garimella 8.0E-09 Yang&and
Yang Garimella
Garimella (2009)
[2009]

6.0E-08 6.0E-09

4.0E-08 4.0E-09

2.0E-08 2.0E-09

0
5.1E-21 6.1E-22
0
0 10 20
20 30
30 40
40 50
50 0 10 20 30 40 50
-2.0E-08 -2.0E-09
Pressure ratio, PR = Pin /Po Pressure ratio, PR = Pin /Po
Pressure ratio, Pin Po Pressure ratio, Pin Po
(a) (b)
6E-10 8E-10
Classical
CNSE NS equations Classical
CNSE NS equations
Extended
ENSE NS equations 7E-10
5E-10 Extended
ENSE NS equations
Mass flow rate, Kg/s

Mass flow rate, Kg/s


Yang
Yang &
Yang and
and Garimella
Garimella
Garimella (2009)
(2009)
[2009] 6E-10 Yang&and
Yang Garimella
Garimella (2009)
[2009]
4E-10 5E-10

3E-10 4E-10

3E-10
2E-10
2E-10
1E-10
1E-10

0 0
0 10 20 30 40 50 0 50 100 150 200

Pressure ratio, PR = Pin Po Pressure ratio, PR = Pin Po


(c) (d)

Figure 5.14 Comparison of the total mass flow rates obtained with the CNSE and the
ENSE with the experimental measurements of Yang and Garimella [2009] as a
function of the pressure ratio PR

One can observe in Table 5.2 that for the case (d), the outlet pressure is less than the
characteristic pressure PC and the diffusive mass flow rate will dominate the flow in
the part of the capillary closer to the outlet. However, it is important to note that even
for cases where the outlet pressure is less than the characteristic pressure PC , the
influence of the diffusive mass flux on the total mass flow rate decreases with increase
in the pressure ratio since the diffusion dominated fraction of the length of the
capillary declines with increasing pressure ratio, as shown in equation (5.45). In
equation (5.45), the term containing lnPin Po  describes the diffusive part of the total
mass flow rate and the other term depicts the convective part. For higher pressure
ratios, the diffusive part of the total mass flow rate becomes negligible and the
difference between the CNSE and the ENSE disappears. Therefore, one does not see
the differences in the mass flow rates predicted by the CNSE and the ENSE even for
the case where the outlet pressure is less than the characteristic pressure PC , as shown
in Figure 5.14(d).

98
1E-09

Classical NS equations
CNSE

1E-10
Extended NS equations
ENSE
Yang and Garimella (2009)
[2009]
Mass flow rate Kg/s

1E-11

1E-12

1E-13

1E-14
1 10 100
Pressure ratio, PR = Pin Po

Figure 5.15 Comparison of the total mass flow rates obtained with the CNSE, the
ENSE and the experimental measurements of Yang and Garimella [2009] as a
function of the pressure ratio PR , [Figure 5.14(d) redrawn]

In order to show the region of the pressure ratio PR where the difference between the
CNSE and the ENSE assumes significance, Figure 5.14(d) is redrawn again in Figure
5.15 as a log-log plot with additional data points with lower values of the pressure ratio.
As one can observe in this figure, the difference between the CNSE and the ENSE is
amplified only below a pressure ratio value of 10. Since all the experimental points of
case (d) of Yang and Garimella (2009) lie above this pressure ratio value, one does not
see any significant influence of the diffusive mass flow rate and the predictions by the
classical and extended equations are identical.

Based on the discussions in this chapter, one can conclude that the characteristics of
gas flows through straight micro-conduits can be accurately predicted with the help of
the analytical solutions of the ENSE, taking the self-diffusive transport of mass into
account. The excellent comparisons of the analytical solutions with the experimental
measurements convincingly prove that the additional mass transport observed in
experiments is nothing but the self-diffusive transport of mass. The entire major
‘microchannel effects’ are explained based on the extended equations with sound
physical reasoning. A characteristic pressure for the microchannel and capillary gas
flows was introduced and the flow characterization based on the comparison of the
local pressure against the characteristic pressure was found to provide excellent
insights into the physics of gas flows through micro-conduits.

99
100
Chapter 6

GAS FLOWS THROUGH COMPLEX


MICROCHANNEL GEOMETRIES
6.1 Introduction

As explained in chapters 4 and 5, ideal gas flows through straight microchannels and
capillaries in the so-called ‘slip-flow’ regime could be solved easily with the help of
the extended Navier-Stokes equations (ENSE) without invoking adhoc assumptions
related to molecule-wall interactions. The excellent agreement of the analytical and
numerical results obtained with the ENSE with the corresponding experimental
measurements has proven that the additional mass flow rate observed in gas flow
experiments in microchannels is indeed the self-diffusion transport of mass caused by
density gradients in the flow. Since the physical reason behind the ‘mystery’ of gas
flows through microchannels has already been answered with the help of the ENSE, it
was felt that more complex problems could be attempted.

The flow conduits in Micro-Electro-Mechanical-Systems (MEMS) may not be simple


straight channels and capillaries. Hence, they often involve complex geometries
including steps and abrupt turns and bends. Therefore, analyzing gas flows, through
complex micro-geometries, is of immense importance and for this reason, the
particular topic has continued to attract the interest of researchers over the years.
Duan and Muzychka [2007, a] had performed theoretical analysis using elliptical
cylindrical coordinates and the separation of variables method to obtain an analytical
solution for gas flows through straight elliptical channels, under flow conditions
belonging to the slip-flow regime. Their prediction employed the slip-flow theory and
was able to predict mass flow rates and pressure distributions in elliptical
microchannels. Duan and Muzychka [2007, b] also obtained a general theoretical
solution in the slip-flow regime for a range of non-circular microchannels involving
regular polygons, trapezoidal and various annular cross-sections based on the aspect
ratio of the geometry. Agrawal et al [2009] simulated gas flows through 90o bends
employing the slip-flow theory and the Direct Simulation Monte Carlo (DSMC)
method.

Simulation of gas flows in backward-facing step geometries is considered to be one of


the benchmark problems to demonstrate the effectiveness of theoretical models and
computational techniques in micro-scale flows. A number of studies have applied the
Lattice Boltzmann Method (LBM) or the Direct Simulation Monte Carlo (DSMC)
method along with the ‘slip-flow’ theory employed in microchannels with a backward-
facing step. Agrawal et al [2005] simulated gas flows in microchannels with a sudden
expansion or contraction in order to obtain insight into flows in complicated
microdevices. They employed the LBM and the computations were performed for
several area and pressure ratios over a range of Knudsen numbers in order to assess
the effects of compressibility and rarefaction. Wu and Lee [2001] demonstrated the
usefulness of the DSMC method in simulating gas flows in backward-facing micro-
step geometries. Kursun and Kapat [2007] also employed the DSMC method to
simulate flows through microchannels with a backward-facing step. The simulations
in this case were conducted with Reynolds numbers of 0.03 - 0.64, Mach numbers of
0.013 - 0.083 and Knudsen numbers of 0.24 - 4.81. In these parameter ranges, they
did not observe any separation region in the flow geometry. They employed an
Information Preservation (IP) method to separate the macroscopic velocity from the
molecular velocity and hence the typical statistical noise generated in DSMC
simulations was minimised.

Figure 6.1 Flow characteristics for the case of Nitrogen flow through microchannel
with a backward-facing step; Kno  0.018 and pressure ratio, PR  2.32 : Celik and Edis
[2006, Figure 6]

Beskok [2001] also employed the DSMC method to simulate high-velocity gas flows
through microchannels with a backward-facing step. Celik and Edis [2006] also
performed similar backward-facing step flows prediction, using the same
microchannel geometry as Beskok [2001], but employing a characteristic-based split
Navier-Stokes finite element solver with second-order slip-velocity and temperature-
jump boundary conditions at the solid walls. The gas flow characteristics as obtained
by Celik and Edis [2006] and Beskok [2001] are shown in Figures 6.1 and 6.2,
102
respectively. It can be observed that the sudden expansion due to the change in the
cross-sectional area results in a temperature drop and hence it cannot be assumed that
the high-velocity gas flows through such microchannels to be isothermal in this case,
unlike the straight microchannels discussed in chapters 4 and 5. Further, the
separation bubble near the bottom wall downstream of the step wall could also be
observed and is shown in Figure 6.1. When the Reynolds number is very low, the
flow is not expected to separate near the bottom wall in microchannel geometries with
backward-facing steps because of the so called ‘slip-velocity’. However, the flow
tends to separate at higher Reynolds number, as is evident in Figure 6.1.

Figure 6.2 Local Mach number and temperature contours for the case of Nitrogen flow
through a microchannel with a backward-facing step; Kno= 0.05, Re=80 : Beskok
[2001, Figure 8]

The above discussion indicates that unlike the solutions obtained in straight
microchannels, it is not possible to use a simple ‘slip-velocity’ model in order to
obtain reliable numerical solutions for gas flows through complex micro-geometries
employing the CNSE. One must resort to special techniques such as the DSMC or the
LBM in order to obtain reasonable flow predictions. As shown later in this chapter, by
employing the ENSE, it is possible to obtain similar accurate flow predictions in such
complex geometries even using commercially available computational fluid dynamics
software.

In this chapter, the results of numerical simulations of gas flows through


microchannels with a backward-facing step, employing the ENSE for the predictions,
are presented and are compared with those obtained by the DSMC and the LBM
simulations available in the literature. As mentioned in chapters 4 and 5, the slip-flow
103
theory, commonly employed in the microchannel literature, was found to lack a sound
physical basis. Further, the analysis of straight microchannels, presented in chapter 5,
clearly indicated that the self-diffusion transport resulted in the additional mass flow
rate observed in the gas flow experiments. The results obtained with the ENSE of gas
flows through microchannels with a backward-facing step substantiate the above
mentioned claims further. Two different backward-facing step cases were treated
numerically: (i) Beskok [2001] and Celik and Edis [2006] and (ii) Chakraborty [2010].
A detailed analysis is provided in the subsequent sections.

6.2 Comparisons of the ENSE solutions of Beskok [2001] and Celik and
Edis [2006]

6.2.1 Geometry and boundary conditions

Beskok [2001] and Celik and Edis [2006] employed identical backward-facing step
geometries in their predictions using the DSMC method and the characteristic-based-
split (CBS) algorithm-based simulations, respectively. Beskok [2001] obtained
numerical discretizations by a spectral element-based compressible Navier-Stokes
algorithm employing the first-order and higher-order slip boundary conditions at the
solid walls. Further, the DSMC method employed the variable hard sphere (VHS)
collision model. Beskok [2001] also reported that while applying the DSMC
algorithm to gas flows through microchannels, slow convergence, large statistical
noise, extensive number of simulated molecules and lack of deterministic surface
effects were encountered. These disadvantages were overcome by using relatively
high-speed gas flows in small aspect ratio geometries with a sufficiently large number
of simulated molecules.

Celik and Edis [2006] employed a modified CBS algorithm to take into account the
slip-velocity and temperature-jump boundary conditions encountered in compressible
flows through micro-sized geometries. The CBS is a general algorithm that can be
employed to solve both compressible and incompressible flow problems. In the CBS
procedure, the numerical instabilities present in the standard Galerkin formulation of
discretization of the governing equation in space are stabilized by discretizing the
equations along the characteristic of the total derivative. Although the solution
procedure employed by Celik and Edis [2006] was different from other conventional
algorithms, the boundary conditions arising out of the slip-flow theory were employed
at the solid boundaries.

The microchannel geometry with a backward-facing step employed by Beskok [2001]


and Celik and Edis [2006] is shown in Figure 6.3 and the salient geometric features of
the channel are summarized in Table 6.1. As can be seen in Figure 6.3, all the linear
dimensions of the channel were non-dimensionalized with the height of the channel at
104
the outlet h, and it was chosen to be1.25μm . When the Knudsen number at the outlet is
considered, its actual value can be obtained by changing either the channel dimensions
or the static pressure. It is generally assumed in the microchannel literature that when
comparing different simulations from geometrically similar channels and if the outlet
Knudsen number and pressure ratio do not change, then the flow characteristics
obtained with various outlet pressure and channel height combinations are identical.
However, since the diffusion velocity is a function of the local pressure and its
gradient [see equation (3.43)] it is possible to obtain different solutions for the same
outlet Knudsen number and pressure ratio combinations by changing the channel
dimensions. Since the exact channel height employed in the studies mentioned in the
literature was not given, the channel dimension at the outlet was chosen to be1.25μm
based on Celik and Edis [2007]. The dimensionless total length L of the channel was
5.6h. Further, the dimensionless step height S was taken to be 0.467. A non-
dimensional approach length l’ of 0.86 was chosen at the beginning of the channel.
The step is located at a dimensionless distance of 1.81 from the inlet of the channel.
The grid divisions at different regions of the geometry are given in Table 6.1.

1.81h
l Horizontal Wall
Inlet
Po  C Symmetry
PS  C1 Outlet
TT  330K Horizontal
Wall h
x2 Vertical
s
Wall
0 x1
Horizontal Wall
x
Wall boundary conditions
Geometry ratios
Horizontal Wall, , , T = 300 K L =x / h= 5.6
l’ =l / h= 0.86
S =s / h= 0.467
Vertical Wall, , , T = 300 K

Figure 6.3 Geometry and boundary conditions employed in the numerical simulations
of a two-dimensional microchannel with a backward-facing step

The ENSE in the total velocity form were employed in all the simulations presented in
this chapter and the same are presented here once again.

105
Continuity equation:
 U iT 
 0 (3.46)
t xi

Momentum equation in the total velocity form:



 U Tj  
 T T
U i U j  
P  ij
 
~ T  U D

j
 g j
  (3.49)
t xi x j x i t

Total energy equation:


 E T   U iT E T  q i 
 PU iT ~T T 
U iD   ij U j  U Tj U Cj
   P   ~ijT   ijT
t x i x i x i x i x i x i x i

 U Tj

 U Dj   g U T (3.76)
j j
t

In equations (3.49) and (3.76), the total molecular momentum transport terms are
given by:
 U Cj U iC  2 C
  
T
ij     ij  U k  m iDU Cj  m DjU iC  2  ij m kDU kC (3.37)
 xi x j  3 xk 3
 
and
2
~ijT   ijC  m iDU Dj   ij m kDU kC (3.67)
3

In equation (3.76), the total energy based on the total kinetic energy E T is given by
1 T
ET  e 
2
 ,
Uj
2
where e is the local internal energy and U Tj is the total velocity.
Since the gas flow considered in this chapter is steady in nature and the influence of
gravity can also be neglected, equations (3.46), (3.49) and (3.76) can be simplified as
given below.

Continuity equation:
U iT 
0 (6.1)
xi

Momentum equation in the total velocity form:


~T
 P  ij
U i U j    
x i
T T

x j xi
(6.2)

106
Total energy equation:
 U iT E T  q i  PU iT  
~T T
U iD   ij U j ~ T
U Tj T
U Cj
  P    ij   ij (6.3)
xi xi xi xi xi xi xi

Table 6.1 Summary of simulation conditions employed by Beskok [2001] and Celik
and Edis [2006] for the case of a microchannel with a backward-facing step

Beskok [2001] / Celik and


Edis [2006]

Data Value No. of grids


employed
Height at outlet, h μm 1.25 110
Dimensionless height at inlet, 0.533 50
Channel 1-S
dimensions Step height, S 0.467 60
Total length of channel, L 5.6 260
Length of approach section, l’ 0.86 40
Inlet Mach number 0.45
Inlet static temperature, TS K 330
Pressure ratio, PR 2.32
Gas used Nitrogen
Molecular diameter, m 3.77  10 10
Outlet Knudsen number, Kno 0.018
Wall temperature, Tw K 300

The following boundary conditions were employed in the simulations as shown in


Figure 6.3:
(i) The outlet pressure of the channel was calculated based on the outlet Knudsen
number Kno= 0.018. Since during the iterative solution procedure the outlet
temperature varied significantly, the average outlet temperature was calculated
frequently during the solution procedure to correct the outlet pressure in order to
obtain the prescribed outlet Knudsen number.
(ii) The inlet static pressure was calculated based on the specified pressure ratio PR
across the channel and the calculated outlet pressure. The dynamic pressure
calculated based on the specified inlet Mach number was added to the static
pressure in order to obtain the total pressure at the inlet.

107
(iii) Similarly, the dynamic temperature was calculated based on the specified inlet
Mach number and was added to the inlet static temperature of 330 K to obtain
the correct thermal boundary conditions at the inlet.
(iv) The symmetry boundary condition was employed along the approach length at
both the top and bottom boundaries.
(v) Since the ENSE were solved in the total velocity form as given in equations
(6.1)-(6.3), the total velocity at the solid boundaries needed to be specified.
Since the convective velocity components satisfy the no-slip boundary
condition at the solid wall, it was sufficient to specify the diffusion velocity as
the total velocity boundary condition at the solid wall. At all the horizontal
walls, the streamwise component of the diffusion velocity was specified, given
  1 P 
byU 1D     , and the normal component of the diffusion velocity was
  P x1 
zero due to the impermeability condition at the walls. Similarly, in the vertical
wall, the cross-stream component of the diffusion velocity was specified, given
  1 P 
byU 2D     , and the streamwise component was set to be zero.
  P x 2 
Furthermore, a constant static temperature of 300 K was specified at the solid
walls.

Simulations were carried out with the commercially available CFD code FLUENT 6.3
and the special boundary conditions, needed for the ENSE computations, were
implemented in the software through User-Defined Functions (UDF). Similarly, the
additional terms in the governing equations were also incorporated as volumetric
source terms. The grid counts in the streamwise and cross stream directions were 200
and 30, respectively.

6.2.2 Comparisons of results with Beskok [2001]

It is important to note that the boundary conditions mentioned by Beskok [2001] are
somewhat ambiguous. The average outlet Knudsen number of the channel Kn , was
calculated based on the following equation.
1
   2 Ma
Kn    (6.4)
 2  Re m T

where  is the specific heat capacity ratio of the gas, Ma the average Mach number at
the outlet and Re m T the Reynolds number of the flow in the channel based on the total
mass flow rate. As stated by Beskok [2001], the flow was maintained subsonic, and
the Mach number was evaluated as an average at the outlet, i.e. Ma  1. Therefore, it is
possible to calculate the maximum possible (or limiting) Reynolds number for a given
outlet Knudsen number and an outlet Mach number of unity, in a given geometry, with
108
the calculations based on equation (6.4). For an outlet Knudsen number of 0.04 (see
Figure 8 of Beskok [2001]), the maximum possible Reynolds number of the channel
was only 37 whereas Beskok [2001] employed a Reynolds number of 80 in the
Beskok simulations. For a Reynolds number of 80, the maximum possible outlet
Knudsen number is only 0.0185 and not 0.04 as mentioned in the paper. The present
author attempted to simulate the flow with the deduced conditions of Beskok [2001]
but they failed to obtain a converged solution. However, since Celik and Edis [2006]
also compared their results with those of Beskok [2001], it was assumed that the
computational conditions, given by Celik and Edis [2006] were identical with the
simulations conditions of Beskok [2001]. Hence, the comparisons of the numerical
simulations of the ENSE were carried out with the flow and boundary conditions
given by Celik and Edis [2006 and 2007].

in Pa

Figure 6.4 Contours of static pressure obtained by employing the ENSE; Kno  0.018 ;
PR  2.32

In Figures 6.4-6.8, the contours of pressure, density, Mach number, static temperature
and Knudsen number, obtained in the numerical simulations of a nitrogen gas flow
through a microchannel with a backward-facing step, employing the ENSE, are
shown, respectively. On comparing them with Figures 6.1 and 6.2, it can be observed
that the profiles obtained by solving the ENSE are almost similar to those obtained in
the simulations of Beskok [2001] and Celik and Edis [2006]. In Figure 6.4, the strong
expansion of the gas can also be observed, near the step wall, due to the sudden
increase in the cross-sectional area.

109
in Kg/m3

Figure 6.5 Contours of density obtained employing the ENSE; Kno  0.018 ; PR  2.32

Figure 6.6 Contours of Mach number, Ma  U 1T T obtained by employing the


ENSE; Kno  0.018 ; PR  2.32

110
in K

Figure 6.7 Contours of static temperature obtained by the ENSE; Kno  0.018 ;
PR  2.32

Figure 6.8 Contours of Knudsen number obtained by employing the ENSE;


Kno  0.018 ; PR  2.32

The local static pressure decreases significantly at the sharp corner and one would
expect that this would result in enhanced diffusion effects. As observed in Figure 6.6,

111
the flow is accelerated rapidly near the sudden expansion of area of the flow geometry
and the flow becomes supersonic. However, the flow becomes gradually subsonic
further downstream of the channel. Towards the exit of the channel, the flow again
accelerates and the velocity increases due to the gradual decrease in pressure. In
Figure 6.6, the recirculation region close to the bottom wall of the channel can be
observed near the step. Further, in Figure 6.7, the drop in the local temperature near
the step caused by the sudden expansion is evident. After a gradual recovery of the
temperature downstream of the step, it decreases again towards the exit of the channel
due to the continued reduction of local pressure values. However, the temperature
near the top and bottom walls does not decrease very much because of lower velocities
and the presence of a re-circulatory region, respectively. The local Knudsen number
profiles are shown in Figure 6.8 and as expected, the Knudsen number increases
towards the exit of the channel. However, the maximum Knudsen number happens to
be at the step because of strong pressure gradients near the step caused by the sudden
expansion of the flow. In general, the profiles of the flow variables, computed with the
ENSE, followed the expected trend and are in good qualitative visual agreement with
the DSMC simulations.

Velocity profiles

In Figure 6.9, the streamwise velocity profiles are plotted as a function of the cross-
stream coordinate at two dimensionless streamwise locations, one in the upstream
direction and the other in the downstream direction of the backward-facing step. The
velocity profiles obtained in the DSMC simulations of Beskok [2001] are also shown
for comparison. It can be observed in Figure 6.9(a) and (b) that the general agreement
between the two profiles obtained by the two methods is very good. However, the
streamwise velocity at the top wall is much larger in the DSMC simulations than in the
present solutions. In other words, the calculated slip-velocity is greater in the DSMC
simulations. Since the outlet Knudsen number is 0.018, the local Knudsen number at
this location is expected to be even lower.

One can observe in Figure 6.8 that the local Knudsen number before the expansion is
only about 0.008. One can calculate, employing equation (4.39) that the diffusion
mass flow rate is only ~0.05% of the convective transport, hence the slip-velocity
must be negligible at this location. Through the ENSE-computations, a near-zero wall
velocity values were obtained near top and bottom walls at this location, unlike the
DSMC solutions of Beskok [2001], which over-estimated the slip-velocity. A detailed
analysis of this feature of the flow is given below.

112
1

Dimensionless cross-stream coordinate


ENSE
DSMC
0.9

0.8

0.7

0.6

0.5

0.4
0 50 100 150 200 250 300 350
Streamwise velocity, U1m/s
[m/s]
(a)
1
ENSE
Dimensionless cross-stream coordinate

0.9 DSMC

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
-50 0 50 100 150 200 250 300 350 400
Streamwise velocity, U1m/s
[m/s]
(b)

Figure 6.9 Comparison of streamwise velocity profiles obtained with the ENSE
(present work) and the DSMC simulations by Beskok [2001]; Kno  0.018 ; PR  2.32 ;
dimensionless axial distance, x h : (a) 1.7 and (b) 2.1

113
35
Top wall,
Top ENSE
wall
30 Bottom wall, ENSE
Bottom wall
3/2l,l,3/2
3/2 Higher
λ order slip
DSMC
25
l,l, Higher
DSMC λ order slip
Wall velocity, m/s

20 O(Kn), 1st order slip


O(Kn)

15

10

-5

-10
1 1.5 2 2.5 3 3.5 4 4.5 5 5.5
Dimensionless streamwise distance

Figure 6.10 Comparison of wall velocity profiles along streamwise coordinate


obtained with the ENSE (present work) and as shown in Beskok [2001], Figure 11;
closed symbols represent values at the top wall whereas open symbols denote values at
the bottom wall; Kno  0.018 ; PR  2.32

In order to further understand this discrepancy in the predicted slip-velocity, the


calculated wall velocity profiles were analysed as a function of the streamwise
coordinate and are shown in Figure 6.10. The continuous line represents the velocity
at the top wall and the dotted line denotes the velocity at the bottom wall. The slip-
velocity values calculated by Beskok [2001], using different wall-slip models, are also
shown for comparison. The solid symbols represent the velocity values at the top wall
and the open symbols denote the values at the bottom wall. The O(Kn) profiles
represent the first-order Maxwell slip model. The l and (3/2)l profiles represent higher
order slip models where the slip information was estimated based on the velocity at a
distance of l and (3/2)l from the wall and l represents the molecular mean free path. It
is interesting to note in Figure 6.10 that the wall velocity calculated by the ENSE is
significantly lower than the values predicted by the various slip-velocity models.
Since the Knudsen number simulated in this flow case is small, it is not possible for
strong self-diffusion effects to be present, hence the wall velocity is expected to be
very low. It was puzzling to note this huge discrepancy, even though the ENSE were
able to calculate all the flow characteristics of straight microchannels and capillaries
successfully, as shown in chapters 4 and 5.

114
Top wall,
Top ENSE
wall
40
Bottom wall, ENSE
Bottom wall
3/2 l,3/2 λ order slip
Higher
l, Higher
λ order slip
30
O(Kn), 1st order slip
O(Kn)
Wall velocity, m/s

20

10

-10
1 1.5 2 2.5 3 3.5 4 4.5 5 5.5

Dimensionless streamwise distance

Figure 6.11 Comparison of wall velocity profiles along streamwise coordinate


obtained with the ENSE (present work) and as shown in Beskok [2001], Figure 11;
closed symbols represent values at the top wall whereas open symbols denote values at
the bottom wall Kno  0.018 ; PR  2.32 ; the top wall is considered to be at 0.9875h and
the bottom wall at 0.01675h

In the solutions obtained by employing the ENSE, the wall velocity values were
estimated ‘exactly at the wall’. However, the slip-velocity values at the wall seemed
to be estimated not exactly at the wall but at small normal distances away from it.
Therefore, it is possible that the wall velocity profiles obtained by the ENSE and the
slip-velocity models, shown in Figure 6.10, may not have been calculated at the same
dimensionless cross-stream location. Further, Beskok [2001] gave the dimensionless
cross-stream coordinates of the top and bottom walls as 0.9875 and 0.01675 instead of
1 and 0, respectively, see Table 2 in Beskok [2001]. Therefore, it is possible that the
slip-velocity values were not estimated at the precise wall locations but some small
normal distance away from the respective walls which might coincide with the first
grid location from the wall.

115
1
ENSE

Dimensionless cross-stream coordinate


DSMC
0.9

0.8

0.7

0.6

0.5

0.4
-14.0 -12.0 -10.0 -8.0 -6.0 -4.0 -2.0 0.0
Cross-stream velocity,,,UU2 [m/s]
m/s

(a)

0.9
Dimensionless cross-stream coordinate

0.8 ENSE
DSMC
0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
-70 -60 -50 -40 -30 -20 -10 0 10
Cross-stream velocity,, UU2 [m/s]
m/s
(b)
Figure 6.12 Comparison of cross-stream velocity profiles obtained with the ENSE
(present work) and the DSMC simulations by Beskok [2001]; Kno  0.018 ; PR  2.32 ;
x h (a) 1.7 and (b) 2.1

116
In order to verify this, the wall velocity profiles shown in Figure 6.10 were redrawn at
the dimensionless cross-stream locations of 0.01675 and 0.9875 to represent the
bottom and top walls, respectively as done by Beskok [2001]. The new profiles were
compared with the profiles predicted with the slip-velocity models given by Beskok
[2001], and are shown in Figure 6.11. The comparison of the two models is
surprisingly good and the wall velocity values predicted by the two methods are
almost identical. Hence it can be concluded that the ‘real’ wall velocity is not very
high at the given Knudsen number but one obtains larger velocity values, as ‘apparent’
slip-velocity, at the neighbourhood of the walls. This is the reason for the discrepancy
in the results presented in Figure 6.10. It can be observed that the discrepancy of the
streamwise velocity at the wall between the ENSE and the DSMC results of Beskok
[2001], as shown in Figure 6.9, also occurs for this reason only. Further, the
dimensionless reattachment location  x h  was found to be 2.8 in the DSMC
simulations by Beskok [2001] and the same value was observed in the simulations
employing the ENSE as seen in Figure 6.11.

ENSE_0.01675 ENSE_0.25 ENSE_0.48325 ENSE_0.75 ENSE_0.9875


Beskok_0.01675 Beskok_0.25 Beskok_0.48325 Beskok_0.75 Beskok_0.9875
1.2

0.8
Local Mach number

0.6

0.4

0.2

-0.2
0 1 2 3 4 5 6
Dimensionless streamwise distance

Figure 6.13 Comparison of local Mach number, U 1T T along the streamwise
distance at different dimensionless cross-stream locations; Main  0.45 ; Kno  0.018 ;
the symbols represent the DSMC data of Beskok [2001] and the lines show the results
obtained with the ENSE (present work)

117
The cross-stream velocity profiles obtained by solving the ENSE at two different
dimensionless streamwise locations are shown in Figure 6.12. The profiles predicted
by the DSMC simulations of Beskok [2001] are also given in this figure. The velocity
profiles, obtained by the two very much different methods, agree exceedingly well as
shown in this figure. Since the impermeability condition at the solid wall needs to be
satisfied, the cross-stream velocity at the solid wall is zero and hence the no-slip
boundary condition is satisfied for the wall normal velocity.

The local Mach number, the streamwise velocity normalized by the local velocity of
sound, profiles as a function of the non-dimensional streamwise coordinate at different
dimensionless cross-stream locations are plotted in Figure 6.13. The Mach number
profiles obtained by the DSMC solutions by Beskok [2001] are also plotted for
comparison. The two sets of profiles agree remarkably well in most regions of the
channel. The wall velocity at the top wall was found to be under-predicted by the
ENSE near the step in comparison to the DSMC simulations.

Pressure profiles

The dimensionless pressure profiles at two streamwise locations are plotted as a


function of the dimensionless cross-stream coordinate in Figure 6.14. The static
1
pressure was non-dimensionalised using the dynamic pressure,
2
 
 in U inT
2
at the inlet.
The pressure profiles obtained from the DSMC simulations by Beskok [2001] are also
shown for comparison along with the profiles obtained by the first and second-order
slip-velocity models.

It can be observed from Figure 6.14(a) that the dimensionless pressure values obtained
by the ENSE are slightly higher than those obtained by the DSMC simulations.
However, the values, predicted by the ENSE, fall within the range predicted by the
slip-velocity models. The profiles obtained by all the three methods agree very well in
Figure 6.14(b). The dimensionless pressure values obtained by the extended equations
at different cross-stream locations of the channel are compared with those predicted by
the DSMC simulations in Figure 6.15. The agreement between the two very different
computational methods is very good.

118
1

Dimensionless cross-stream coordinate ENSE


ENSE
DSMC
DSMC
0.9
Lamda
l, Higher order slip
O(Kn), 1st order slip
O(Kn)

0.8

0.7

0.6

0.5

0.4
3.7 3.8 3.9 4.0 4.1 4.2 4.3 4.4
Dimensionless pressure
(a)

0.9 ENSE
ENSE
Dimensionless cross-stream coordinate

DSMC
DSMC
0.8 l, Higher order slip
Lamda
O(Kn), 1st order slip
O(Kn)
0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
2.8 2.9 3.0 3.1 3.2 3.3 3.4
Dimensionless pressure
(b)

Figure 6.14 Comparison of dimensionless static pressure, P 0 .5  in U inT  profiles  2



obtained with the ENSE (present work) and the DSMC simulations by Beskok [2001];
slip-velocity predictions given by Beskok [2001] are also shown; Kno  0.018 ;
PR  2.32 ; x h (a) 1.7 and (b) 2.1
119
ENSE_0.01675 ENSE_0.25 ENSE_0.48325 ENSE_0.75 ENSE_0.9875
Beskok_0.01675 Beskok_0.25 Beskok_0.48325 Beskok_0.75 Beskok_0.9875
8

7
Dimensionless pressure

0
0 1 2 3 4 5 6
Dimensionless streamwise distance

Figure 6.15 Comparison of dimensionless static pressure, P 0 .5  in U inT  along the  2



streamwise distance at different dimensionless cross-stream locations; Main  0.45 ;
Kno  0.018 ; the symbols represent the DSMC data of Beskok [2001] and the lines
show the results obtained with the ENSE

Temperature profiles

The static temperature profiles obtained with the ENSE are shown at two
dimensionless streamwise locations in Figure 6.16. The temperature profiles obtained
by the DSMC simulations and the slip-velocity models are also shown. It can be
observed in Figure 6.16(a) that the temperature values predicted by the extended
equations are much lower than those given by the DSMC simulations. Due to the
expansion caused by the sudden increase in the area, the thermal energy is converted
into kinetic energy and hence, the temperature was significantly reduced. In the
absence of local measurements, it will be difficult to conclude which method predicted
the temperature profiles accurately.

120
1
ENSE

Dimensionless cross-stream coordinate


DSMC
0.9 Lamda
l
O(Kn)

0.8

0.7

0.6

0.5

0.4
270 275 280 285 290 295 300 305
Static temperature, K

(a)

1
Dimensionless cross-stream coordinate

0.9

0.8 ENSE
DSMC
0.7 Lamda
l
0.6 O(Kn)

0.5

0.4

0.3

0.2

0.1

0
220 230 240 250 260 270 280 290 300 310 320

Static temperature, K
(b)
Figure 6.16 Comparison of static temperature profiles obtained with the ENSE
(present work) and the DSMC simulations by Beskok [2001]; Kno  0.018 ;   2.32 ;
x h (a) 1.7 (b) 2.1

121
It can be observed from the above mentioned discussions that the general flow features
of gas flows through a microchannel with backward facing step is well captured by the
extended Navier-Stokes equations. It is necessary to mention here that the excellent
agreement of the profiles of various flow variables indicates that the outlet Knudsen
number for the DSMC case mentioned in Beskok [2001] is indeed 0.018 only.

6.2.3 Comparisons of results with Celik and Edis [2006]

ENSE_0.01456 ENSE_0.2476 ENSE_0.4978 ENSE_0.755


ENSE_0.9901 C&D_0.01456 C&D_0.2476 C&D_0.4978

1.2

0.8
Local Mach number

0.6

0.4

0.2

-0.2
0 1 2 3 4 5 6
Dimensionless streamwise distance

Figure 6.17 Comparison of local Mach number, U 1T T along the streamwise
distance at different dimensionless cross-stream locations; Main  0.45 ; Kno  0.018 ;
the symbols represent the simulation data of Celik and Edis [2006] and the lines depict
the results obtained with the ENSE

Celik and Edis [2006] employed the slip-velocity model in the CBS scheme in order to
provide a solution for gas flows through microchannels with a backward-facing step.
In Figure 6.17, the local Mach number values obtained with the extended equations
are compared with those calculated based on the CBS algorithm by Celik and Edis
[2006]. The agreement between the two profiles is good. There is a discrepancy in
the profiles close to the top wall in the neighbourhood of the step because of the
differences in the calculated wall velocity, as described in Figures 6.10 and 6.11. The
agreement between the profiles obtained by the two methods towards the outlet of the
channel is also very good.

122
In Figure 6.18, the dimensionless pressure profile obtained with the ENSE along the
streamwise direction at the middle section of the channel is shown. The profiles
obtained by the simulations employing the slip-velocity theory carried out by Celik
and Edis [2006], Beskok [2001 (b)] and Baysal and Aslan [2002] are also depicted.
The profiles agree with one another exceedingly well throughout the entire length of
the channel. The small discrepancy observed near the inlet could be attributed to the
minor variations in the inlet conditions since the exact boundary conditions used in the
simulations of Celik and Edis [2006], Beskok {2001(b)] and Baysal and Aslan [2002]
were not mentioned in the papers.

7 ENSE

Celik_Edis [2006]
pressure

6
Beskok [2001(b)]
Mach number

5
Baysal & Aslan [2002]
Dimensionless

3
Local

0
0 1 2 3 4 5 6
Dimensionless streamwise distance


Figure 6.18 Comparison of dimensionless pressure, P 0 .5  in U inT  along the
2

streamwise distance at the middle section of the channel; Main  0.45 ; Kno  0.018 ;
the symbols represent the simulation data of Celik and Edis [2006], Beskok [2001 (b)]
and Baysal and Aslan [2002] and the lines depict the results obtained with the ENSE

Further, the local Knudsen number profile obtained with the ENSE along the
streamwise direction at the middle section of the channel is compared with the profiles
obtained by Celik and Edis [2006] and Baysal and Aslan [2002] in Figure 6.19. The
agreement between the profiles is excellent throughout the length of the channel
except in the region close to the step. As observed in Figures 6.14 and 6.16, the static
pressure and temperature profiles obtained by the ENSE and the slip-velocity based
theory did not match in the neighbourhood of the step. It is felt that in the absence of
accurate experimental measurements, it is not possible to ascertain the validity of the
profiles obtained by either the ENSE or the DSMC simulations or any other method.
It may be argued that unlike other methods employed in the simulations of micro-gas
flows, the ENSE do not have empirical assumptions regarding the slip-velocity, hence
123
the possibility of the profiles obtained with these equations being correct is high. It is
suggested that this aspect of the gas flow through backward-facing step geometries
needs to be investigated comprehensively in further studies of this kind of flows.

0.02

0.016
Local Knudsen number

0.012

ENSE
0.008 Celik and Edis [2006]
Baysal & Aslan [2002]

0.004

0
0 1 2 3 4 5 6
Dimensionless streamwise coordinate

Figure 6.19 Comparison of local Knudsen number, Kn  l h along the streamwise


direction at the middle of the channel; Main  0.45 ; Kno  0.018 ; the symbols
represent the simulation data of Celik and Edis [2006] and Baysal and Aslan [2002]
and the lines depict those obtained with the ENSE

Based on the various discussions mentioned in this section, it can be concluded that
the ENSE were successfully employed in accurately predicting all the characteristics
of gas flows through microchannels with a backward-facing step. Based on the
comparisons with the DSMC and slip-velocity theory based simulations, it was
evident that the ENSE can be employed satisfactorily in the Knudsen number range
studied, to predict micro-channel gas flows with complex flow geometries.

6.3 Comparisons of the ENSE solutions with Chakraborty [2010]

The unpublished data of the DSMC simulations of a helium gas flow through a
microchannel with a backward-facing step were obtained from Chakraborty [2010]
(personal communication). Simulations were performed with the ENSE and
comparisons were made with the profiles obtained in the DSMC simulations, as
explained in this section.

124
6.3.1 Geometry and boundary conditions

The geometry and boundary conditions employed in the simulations mentioned in this
section is shown in Figure 6.20. The linear dimensions of the channel were non-
dimensionalized with the half-height of the channel a, which is equal to 0.5μm . The
total length of the channel was taken to be 10a and the height of the channel at the exit
was 2a. At the inlet section, the channel height was equal to a. The step was located
at a distance of 3a from the entrance. The grid divisions at different regions of the
channel are given in Table 6.2.

The outlet Knudsen number Kno was varied in two steps; (i) 0.01 and (ii) 1. Helium
was employed as the working fluid and the simulation conditions and the properties of
the gas used are also given in Table 6.2. The static temperature of the flow at the inlet
was 300 K and all the solid walls were also maintained at the same temperature. The
outlet pressure of the channel was calculated based on the specified Knudsen number
and the height of the channel at the outlet. Based on the specified pressure ratio PR ,
the inlet static pressure at the inlet was calculated and the same was maintained at the
inlet.

Wall

Inlet
3a
Wall Outlet 2a

y a Wall
Wall
x
10a = 5 m

Wall boundary conditions

Horizontal wall: , ,

Vertical wall: ;

Figure 6.20 Schematic representation of the geometry and boundary conditions


employed in the simulations of the microchannel with a backward-facing step of
Chakraborty [2010]

125
6.3.2 Results and Discussion for the case of Kno=0.01

Comparisons of pressure profiles


The pressure distributions obtained with the ENSE along the streamwise direction at
different dimensionless cross-stream locations are plotted in Figures 6.21 - 6.23. The
pressure profiles near the bottom wall, at the middle of the channel and at the top wall
are shown. The local static pressure was non-dimensionalised using the outlet
pressure of the channel. The dimensionless pressure values obtained in the DSMC
simulations of Chakraborty [2010] are also shown for comparison. It can be observed
that there is excellent agreement between the profiles obtained by the two methods
near the inlet and outlet boundaries.

Table 6.2 Summary of simulation conditions employed by Chakraborty [2010]


Chakraborty[2010]
Data
Value No. of grid employed
Height at outlet, 2a μm 1.0 48
Height at inlet, a μm 0.5 24
Channel
Step height, a μm 0.5 24
dimensions
Total length of the channel, 10a μm 5.0 210
Length of approach section, 3a μm 1.5 60
Inlet static temperature, TS K 300
Outlet Knudsen number, Kno 0.01 and 1.0
Kno  0.01 1.75
Pressure ratio, PR
Kno  1.0 1.9
Gas used Helium
Wall temperature, Tw K 300
Atomic diameter, m 2.3 x 10-10
Dynamic viscosity, Pa.s 2.066 x 10-5

However, there are discrepancies in the pressure profiles in the neighbourhood of the
backward-facing step. In the neighbourhood of the step, the pressure is reduced
drastically because of the change in the cross-sectional area of the channel and part of
the pressure is converted into kinetic energy of the fluid. Subsequently, there is a
pressure recovery towards the outlet. The pressure profile predicted with the ENSE
depicts a sharp fall in pressure near the step compared with the DSMC simulations. As
mentioned in the previous section, the characteristics of expansion near the step need
to be investigated thoroughly with detailed experimental verifications.
126
1.3

1.2

Nondimensionalised pressure
1.1

1.0

0.9 ENSE

0.8 DSMC

0.7

0.6

0.5

0.4
0 1 2 3 4 5 6 7 8 9 10
Dimensionless streamwise coordinate

Figure 6.21 Comparison of dimensionless pressure P Po  profiles along the length of


the channel obtained with the ENSE and the DSMC at the dimensionless cross-stream
coordinate (1/30)a; Kno  0.01

2.0

1.8
ENSE
Nondimensionalised pressure

1.6
DSMC
1.4

1.2

1.0

0.8

0.6

0.4

0.2

0.0
0 1 2 3 4 5 6 7 8 9 10
Dimensionless streamwise coordinate

Figure 6.22 Comparison of dimensionless pressure P Po  profiles along the length of


the channel obtained with the ENSE and the DSMC at the dimensionless cross-stream
coordinate a; Kno  0.01

127
1.8

1.6 ENSE
Nondimensionalised pressure 1.4
DSMC

1.2

1.0

0.8

0.6

0.4

0.2

0.0
0 1 2 3 4 5 6 7 8 9 10
Dimensionless streamwise coordinate

Figure 6.23 Comparison of dimensionless pressure P Po  profiles along the length of


the channel obtained with the ENSE and the DSMC at the dimensionless cross-stream
coordinate 2a; Kno  0.01

1.8
Dimensionless cross-stream coordinate

1.6

1.4

1.2

0.8 X=2a
X=4a
0.6
X=6a
0.4
X=8a
0.2 X=10a
0
0.8 0.9 1 1.1 1.2 1.3 1.4
Dimensionless pressure

Figure 6.24 Dimensionless pressure P Po  profiles obtained with the ENSE at different
streamwise locations; Kno  0.01

128
2

Dimensionless cross-stream coordinate


1.8

1.6

1.4 ENSE_X=4a
ENSE_X=10a
1.2
DSMC_X=4a
1 DSMC_X=10a
0.8

0.6

0.4

0.2

0
0.6 0.7 0.8 0.9 1 1.1 1.2 1.3
Dimensionless pressure

Figure 6.25 Comparison of dimensionless pressure P Po  profiles at two streamwise


locations obtained with the ENSE and the DSMC; X = 4a and 10a; Kno  0.01

In Figure 6.24, the dimensionless pressure profiles at different streamwise locations


are plotted and one can observe a reduction in pressure near the step and subsequent
recovery towards the exit of the channel. Further, in Figure 6.25, the dimensionless
pressure profiles obtained with the ENSE are compared with the DSMC simulations at
two streamwise positions. There are significant deviations between the two pressure
profiles at the streamwise position 4a. Since this position is very close to the
backward-facing step, the deviation is significant throughout the height of the channel.
However, at the streamwise position 10a, both the profiles converge to a single value,
as expected. It can also be seen that the pressure values obtained by the DSMC
simulations at the outlet of the channel are not the same and only the average pressure
is equal to unity. However, the pressure is constant throughout the height of the
channel in the simulations with the ENSE. One can easily deduce from these profiles
that the diffusion mass flow rate, as given by equation (3.9), at the boundary and also
at different parts of the channel will be different in the two methods and hence the
profiles of flow variables can differ significantly.

It is important to note that although the same average inlet and outlet pressure values
were specified in the simulations with the ENSE and in the DSMC simulations, the
implementations of the boundary conditions in the two methods are not the same.
Therefore, it is possible that there may be significant deviations in the results obtained
from the two methods. Hence it is necessary not only to specify the average values of
the flow variables at the boundaries but also to mention the way in which the boundary
conditions were implemented in the solution procedure.

129
Velocity profiles

The dimensionless streamwise velocity profiles along the channel at different cross-
stream locations, namely (1/30)a, a and 2a, are plotted in Figures 6.26-6.28,
respectively. The streamwise velocity was non-dimensionalised using the average
velocity at the outlet of the channel. There is huge discrepancy between the velocity
values predicted by the two methods, especially the profiles at the wall. In Figure
6.27, the velocity profiles are plotted in the middle of the channel, i.e., the
dimensionless cross-stream distance a. It can be observed that for the streamwise
coordinate X < 3, the wall velocity predicted by the ENSE is almost zero whereas
much higher values were obtained from the DSMC simulations. However, the
predictions of the ENSE improve for the streamwise coordinate X > 3 which
represents the middle of the channel away from the wall. Further, one can also
observe in Figure 6.28 that the discrepancy between the predictions of the two
methods is large.

0.5
Dimensionless streamwise velocity

ENSE_(1/30)a
0.4 DSMC_(1/30)a

0.3

0.2

0.1

0.0

-0.1
2 3 4 5 6 7 8 9 10

Dimensionless streamwise coordinate

Figure 6.26 Comparison of dimensionless streamwise velocity U 1T U oT  profiles along


the length of the channel obtained with the ENSE and the DSMC at the dimensionless
cross-stream coordinate (1/30)a; Kno  0.01

In the case of the ENSE, the wall velocity is always calculated exactly ‘at the wall’. In
the present case under consideration, the outlet Knudsen number is only 0.01.
Therefore, the self-diffusion transport of mass is expected to be very small in
comparison with the convective mass transport, hence the calculated wall velocity by
the ENSE is lower. It has already been discussed for Figures 6.10 and 6.11 that the
wall velocity might not have been calculated ‘exactly at the wall’ in the case of the
130
DSMC simulations. This argument seems to be validated by the fact that the velocity
profile at the bottom wall was not calculated at the dimensionless cross-stream
location of zero but at (1/30)a, a small distance away from the bottom wall, as shown
in Figure 6.26.

2.0

1.8
Dimensionless streamwise velocity

1.6

1.4

1.2

1.0
ENSE_X=a
0.8
ENSE_X=a+(1/30)a
0.6 DSMC_X=a
0.4

0.2

0.0
0 1 2 3 4 5 6 7 8 9 10
Dimensionless streamwise coordinate
Figure 6.27 Comparison of dimensionless streamwise velocity U 1T U oT profiles along
the length of the channel obtained with the ENSE and the DSMC at the dimensionless
cross-stream coordinates a and a+(1/30)a; Kno  0.01

In order to evaluate this argument further, the streamwise velocity profiles along the
length of the channel are also plotted at the cross-stream locations (1/30)a, a+(1/30)a
and 2a-(1/30)a in Figures 6.26-6.28, respectively. It is clear that the predictions of
the ENSE have improved significantly. It is important to note that the outlet Knudsen
number of 0.01 is only the beginning of the so-called ‘slip-flow’ regime, hence the
slip-velocity is not expected to be very high. Therefore, it is impossible to assume the
wall velocity values predicted by the DSMC simulations as the ‘slip-velocity’.
However, it is possible that it can be misinterpreted as the ‘slip-velocity’ and the
reader could be misled. Further, it is also possible to miscalculate the molecular
momentum transport to the wall if the velocity profiles were assumed to be the wall
velocity. Interestingly, the solutions obtained by the ENSE pose no such discrepancy
in the understanding of the flow characteristics. The calculated wall velocity is
consistent with the calculations presented in section 4.2.3 and, more importantly, all
the flow characteristics are also determined accurately.

131
0.6

ENSE_X=2a
Dimensionless streamwise velocity
0.5
ENSE_X=2a-(1/30)a
DSMC_X=2a
0.4

0.3

0.2

0.1

0.0

-0.1
0 1 2 3 4 5 6 7 8 9 10
Dimensionless streamwise coordinate

Figure 6.28 Comparison of dimensionless streamwise velocity U 1T U oT  profiles along


the length of the channel obtained with the ENSE and the DSMC at dimensionless
cross-stream coordinates 2a and 2a-(1/30)a; Kno  0.01

2
Dimensionless cross-stream coordinate

1.8

1.6

1.4

1.2

0.8 X=2a

0.6 X=4a
X=6a
0.4
X=8a
0.2
X=10a
0
-0.5 0 0.5 1 1.5 2 2.5
Dimensionless streamwise velocity

Figure 6.29 Dimensionless streamwise velocity U 1T U oT  profiles obtained with the


ENSE at different cross-stream locations; Kno  0.01

132
2

Dimensionless cross-stream coordinate 1.8

1.6

1.4

1.2

0.8

0.6 ENSE
DSMC
0.4

0.2

0
-0.5 0 0.5 1 1.5 2 2.5
Dimensionless streamwise velocity

Figure 6.30 Comparison of dimensionless streamwise velocity U 1T U oT  profiles


obtained with the ENSE and the DSMC at the streamwise location 4a; Kno  0.01

1.8 ENSE
Dimensionless cross-stream coordinate

DSMC
1.6

1.4

1.2

0.8

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
Dimensionless streamwise velocity

Figure 6.31 Comparison of dimensionless streamwise velocity U 1T U oT  profiles


obtained with the ENSE and the DSMC at the streamwise location 10a; Kno  0.01

133
In Figure 6.29, the streamwise velocity profiles, obtained with the ENSE at different
cross-stream locations are plotted and the flow separation at the bottom wall can be
seen. However, this was not present in the profiles obtained from the DSMC
simulations. In Figures 6.30 and 6.31, the streamwise velocity profiles obtained with
the ENSE are compared with those calculated from the DSMC simulations. In Figure
6.30, it can be seen that because of the presence of the separated flow regime close to
the bottom wall, the spread in the velocity profile in the cross-stream direction is
limited in the case of the ENSE. Therefore, the maximum velocity obtained by the
ENSE is higher than that of the DSMC solutions. In Figure 6.31, the velocity profile
obtained with the extended equations is more like a parabolic profile with negligible
wall velocity. In contrast, the profile obtained with the DSMC solutions is flat with a
high wall velocity. As stated earlier, such high wall velocity values are not possible
for the Knudsen number range considered in this study and this can affect the
calculations of the molecular momentum transfer to the wall.

6.3.3 Results and discussion of the case Kno  1

During the simulations of this case with very high Knudsen number, it was observed
that the convergence of the solution was difficult. It can be observed from equations
(6.1)–(6.3) that while implementing the ENSE in the commercially available CFD
software, a number of terms had to be incorporated in the solution procedure apart
from specifying the tangential velocity at the solid walls. Since it was not possible to
treat these additional terms as fluxes through the control volume faces in the
commercial CFD software employed, they were converted into volumetric source
terms and incorporated into the governing equations. When the self-diffusion mass
transport is small compared with the convective transport, the above-mentioned
procedure can be acceptable and may not involve any computational difficulty.
However, when the Knudsen number is high, as in the present case, the self-diffusion
mass transport is greater than the convective transport and the above-mentioned
simplification may create convergence difficulties, especially near the outlet boundary
where the Knudsen number is the maximum. Even a small error in the pressure
calculations can drastically change the self-diffusion of mass because of lower static
pressure values at higher Knudsen numbers. Because of this difficulty with the chosen
methodology, it was difficult to obtain converged solutions. It is suggested to treat the
additional terms as flux terms by developing a proper computational fluid dynamics
tool for the ENSE.

Pressure profiles

In Figures 6.32–6.34, the dimensionless pressure profiles obtained with the ENSE are
plotted along the length of the channel at different cross-stream locations. As
mentioned earlier, the local static pressure values were non-dimensionalised using the
outlet pressure of the channel. Further, the dimensionless pressure values calculated

134
from the DSMC simulations are also shown for comparison. It can be observed that
the two profiles agree very well with each other throughout the channel.

4.0
ENSE
3.5 DSMC
Dimensionless pressure

3.0

2.5

2.0

1.5

1.0

0.5
0 1 2 3 4 5 6 7 8 9 10
Dimensionless streamwise distance

Figure 6.32 Comparison of dimensionless pressure P Po  profiles obtained with the


ENSE and the DSMC along the length of the channel at the cross-stream location
(1/30)a; Kno  1

2.0

ENSE
1.8
DSMC
Dimensionless pressure

1.6

1.4

1.2

1.0

0.8
0 1 2 3 4 5 6 7 8 9 10
Dimensionless streamwise distance

Figure 6.33 Comparison of dimensionless pressure P Po  profiles obtained with the


ENSE and the DSMC along the length of the channel at the cross-stream location a;
Kno  1
135
2.0

1.9
ENSE
1.8
Dimensionless pressure

DSMC
1.7

1.6

1.5

1.4

1.3

1.2

1.1

1.0
0 1 2 3 4 5 6 7 8 9 10
Dimensionless streamwise distance

Figure 6.34 Comparison of dimensionless pressure P Po  profiles obtained with the


ENSE and the DSMC along the length of the channel at the cross-stream location 2a;
Kno  1

2
Dimensionless cross-stream coordinate

1.8

1.6

1.4

1.2

0.8
X=2a
0.6 X=4a
X=6a
0.4
X=8a
0.2 X=10a
0
0.8 0.9 1 1.1 1.2 1.3 1.4 1.5 1.6 1.7
Dimensionless pressure

Figure 6.35 Dimensionless pressure P Po  profiles obtained with the ENSE at


different streamwise locations; Kno  1

136
It can be seen in Figure 6.32 that the pressure values obtained by the DSMC
simulations at the junction between the step wall and the bottom wall are not feasible.
As observed in Figure 6.37, no flow separation was observed near the bottom wall of
the channel in the DSMC simulations. Hence there was no possibility of stagnating
flow at the step wall, which could have resulted in a significant pressure increase. In
the absence of any such stagnating flow conditions, the pressure cannot increase
drastically as shown in Figure 6.32. It is not possible to specify any reason for this
kind of increase in the pressure profile. Apart from this one discrepancy, the pressure
profiles are almost identical in both cases.

2
Dimensionless cross-stream coordinate

1.8

1.6 ENSE_X=4a

1.4 ENSE_X=10a
DSMC_X=4a
1.2
DSMC_X=10a
1

0.8

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Dimensionless pressure

Figure 6.36 Comparison of dimensionless pressure P Po  profiles obtained with the


ENSE and the DSMC at two streamwise locations; X = 4a and 10a; Kno  1

In Figure 6.35, the dimensionless pressure profiles obtained by the ENSE are shown at
different streamwise locations. The pressure remains almost constant in the cross-
stream direction throughout the channel and the flow behaves like a fully developed
flow in most regions of the channel, i.e. the cross-stream velocity will be very small.
However, the cross-stream velocity is significant near the expansion region where the
flow diffuses into the larger area of cross-section. In Figure 6.36, the dimensionless
pressure profiles obtained with the ENSE and the DSMC simulations are compared at
two streamwise locations. The pressure profiles obtained from the two methods agree
well in both locations.

137
Velocity profiles

8
ENSE
Dimensionless streamwise velocity
7 DSMC

-1
0 1 2 3 4 5 6 7 8 9 10
Dimensionless streamwise distance from the inlet
Figure 6.37 Comparison of dimensionless streamwise velocity U 1T U oT  profiles
obtained with the ENSE and the DSMC along the length of the channel at the cross-
stream location (1/30)a; Kno  1

2.0
ENSE
1.8
DSMC
Dimensionless streamwise velocity

1.6

1.4

1.2

1.0

0.8

0.6

0.4

0.2

0.0
0 1 2 3 4 5 6 7 8 9 10
Dimensionless streamwise distance from the inlet
Figure 6.38 Comparison of dimensionless streamwise velocity U 1T U oT  profiles
obtained with the ENSE and the DSMC along the length of the channel at the cross-
stream location a; Kno  1

138
1.4
ENSE
1.3
Dimensionless streamwise velocity
DSMC
1.2

1.1

1.0

0.9

0.8

0.7

0.6

0.5

0.4
0 1 2 3 4 5 6 7 8 9 10
Dimensionless streamwise distance from the inlet

Figure 6.39 Comparison of dimensionless streamwise velocity U 1T U oT  profiles


obtained with the ENSE and the DSMC along the length of the channel at the cross-
stream location 2a; Kno  1

1.8
Dimensionless cross-stream coordinate

1.6

1.4

1.2
X=2a
1 X=4a
0.8 X=6a
X=8a
0.6
X=10a
0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
Dimensionless streamwise velocity

Figure 6.40 Dimensionless streamwise velocity U 1T U oT  profiles obtained with the


ENSE and the DSMC at different cross-stream locations; Kno  1

139
The streamwise velocity profiles along the channel obtained with the ENSE are
compared with those calculated from the DSMC simulations in Figures 6.37–6.39.
The agreement between the profiles obtained by the two methods is very good.
Similarly to Figure 6.32, huge velocity values are observed near the junction of the
bottom and step walls in Figure 6.37. This velocity magnitude appears infeasible and
artificial and no physically plausible reason can be given for this behaviour of the
streamwise velocity. Further, a sharp discontinuity was observed near the step in the
profiles obtained with the ENSE, as seen in Figure 6.38. It is considered that this was
also because of the additional terms added as volumetric source terms instead of
surface fluxes in the solution procedure. However, this could not be verified because
of the unavailability of a suitable computational fluid dynamics code to treat the ENSE
properly. The agreement between the solutions obtained by the ENSE and the DSMC
solutions at the top wall is very good, as can be observed in Figure 6.39.

2
Dimensionless cross-stream coordinate

1.8

1.6

1.4

1.2

0.8
ENSE
0.6
DSMC
0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Dimensionless streamwise velocity

Figure 6.41 Comparison of dimensionless streamwise velocity U 1T U oT  profiles


obtained with the ENSE and the DSMC at the streamwise location 4a; Kno  1

In Figure 6.40, the dimensionless streamwise velocity profiles obtained with the ENSE
at different streamwise locations are shown and the significant wall velocity values in
these profiles can be observed. In Figure 6.29, the velocity values at the wall were
almost negligible because the Knudsen number of the flow was very small. In
contrast, since the outlet Knudsen number considered in this flow case is unity, the
self-diffusion mass transport is dominant in comparison with the convective transport
of mass and hence one can observe high wall velocity values in the solutions obtained
by the ENSE. Further, it is interesting to note in Figure 6.40 that the velocity gradients
at the walls are relatively smaller than those in Figure 6.29. Therefore, the velocity

140
values at the wall are not different from those at a location slightly away from it,
unlike the case with Kno  0.01 . Therefore, it was possible to obtain good agreement
between the wall velocities exactly estimated ‘at the wall’ and the DSMC simulations
in Figures 6.37–6.39 unlike the profiles shown in Figures 6.26–6.28.

2
ENSE
Dimensionless cross-stream coordinate

1.8
DSMC
1.6

1.4

1.2

0.8

0.6

0.4

0.2

0
0.6 0.7 0.8 0.9 1 1.1 1.2
Dimensionless streamwise velocity

Figure 6.42 Comparison of dimensionless streamwise velocity U 1T U oT  profiles


obtained with the ENSE and the DSMC at the streamwise location 10a; Kno  1

In Figures 6.41 and 6.42, the dimensionless streamwise velocity profiles are plotted at
two different streamwise locations. It can be observed in Figure 6.41 that the profiles
do not agree very well in the bottom half of the geometry and the agreement in the top
half is better. The top and bottom wall velocity values are also different in both cases.
Since this position is in the immediate downstream of the backward-facing step, the
characteristics of the expansion from the narrow to large area determine the velocity
profiles at this location. The small discrepancies in the pressure profiles obtained by
the two methods, as observed in Figure 6.36, could be the reason for the deviations in
the velocity profiles depicted in Figure 6.41. However, the streamwise velocity
profiles obtained with the ENSE and the DSMC simulations agree very well at the
outlet plane, as shown in Figure 6.42, where the pressure profiles were also almost
identical, as shown in Figure 6.36.

As described in this chapter, the ENSE were employed successfully to predict the
characteristics of ideal gas flows through microchannels with a backward-facing step.
In this chapter, simulations were carried out with a wide range of Knudsen numbers
and the extended equations were able to predict the flow characteristics accurately
141
even at an outlet Knudsen number as high as unity. It is well known that even the
continuum-based slip-flow theory cannot predict the flow characteristics of
microchannels with outlet Knudsen numbers of the order of unity. This flow regime
of Kn  1is called the transition regime or moderately rarefied flow regime where the
continuum hypothesis is not expected to be valid. It must be stressed that the extended
equations are based on the continuum assumption and the characteristics of flows even
beyond the slip-flow regime, i.e. Kno  0.01, could be determined accurately. This
raises serious questions about the selection of the characteristic dimension in the
definition of the Knudsen number, as outlined in section 4.5. Based on the analysis
mentioned in this chapter, it is evident that if all the details of the boundary conditions
employed in experiments and simulations are made available, then the predictions can
be more accurate. It is also important to state here that most of the additional terms in
the extended equations have been added as volumetric source terms in the simulations
presented in this thesis since the available computational fluid dynamics tools have
been developed for the classical Navier-Stokes equations. The volumetric source
terms thus added have been found to cause stability and oscillations under certain flow
conditions and obtaining converged solutions was difficult. It is stressed that
developing suitable computational fluid dynamics tools is very much necessary for
employing these valuable equations with ease.

142
Chapter 7

SOLUTIONS TO PROBLEMS WITH STRONG


TEMPERATURE GRADIENTS: SHOCK WAVES

7.1 Introduction

In the earlier chapters, the ENSE proposed in this thesis, were successfully applied to
solve isothermal problems related to gas flows through microchannel flows with
excellent agreement with the corresponding experimental measurements. Though some
non-isothermal cases in gas flows through microchannels with a backward facing step
were treated in chapter 6, it did not involve very strong temperature gradients. As
observed in equation 3.9, the self-diffusion transport of mass is governed by the
presence of both density/pressure and temperature gradients. Hence, it is required to
employ the ENSE in solving problems involving strong temperature gradients in order
to completely validate the derivations presented in chapter 3.

Numerical prediction of the characteristics of normal shock waves is one such problem
where both strong pressure and temperature gradients are present. As observed in
Figure 1.4, the CNSE do not predict the shock structure problem, which deals with the
spatial variation of gas properties across a stationary, planar, one dimensional shock
even in a monoatomic gas accurately. Generally, shocks are often considered to be
discontinuities where the gas properties vary abruptly within a short distance of the
order of few molecular mean free paths. In the numerical analysis, the properties are
traditionally modelled with jump conditions defined based on certain thermodynamic
relationships, famously known as Rankine-Hugoniot relationships. However, as
observed in the experimental measurements, the physical properties of the gas across
the shock does not change abruptly but in fact, vary continuously from their upstream
to downstream values over a distance of few molecular mean free paths; see Alsmeyer
[1976]. Further, gas flows within the shock layer are considered to be far from local
thermodynamic equilibrium and have Knudsen numbers in the intermediate range,
typically Kn=0.2~0.3; see Greenshields and Reese [2007(b)]. These are the reasons
stated, in general, for the classical equations not predicting the shock structures
accurately. Hence, the shock structure problem was considered to be particularly
appropriate for testing various extended hydrodynamic models, such as Burnett
equations, see Balakrishnan (2004).

Further, this problem also has few more favourable conditions to test the
hydrodynamic models, such as,
a. there are no solid boundaries to consider and hence one need not worry whether
to apply slip or no-slip boundary condition,
b. the upstream and downstream boundary conditions are clearly defined through
the Rankine–Hugoniot relations, with all gradients of flow quantities tending to
zero far upstream and downstream of the shock and
c. the problem is one-dimensional and steady.

The theoretical evaluation of hydrodynamic models is generally carried out with


monoatomic gases because a monoatomic gas possesses only translational degrees of
freedom and no modes of vibration or rotational degrees of freedom. Further, it
cannot dissociate and only ionizes at higher temperatures and therefore its
thermodynamic behaviour is generally much simpler than that of polyatomic gases.

Dimensionless distance through the shock

Figure 7.1 Comparison of shock structures obtained by the CNSE and the MBNSE
with the experimental measurements of Alsmeyer [1976], as calculated by
Greenshields and Reese [2007(b)]; Inlet Mach number = 9

Greenshields and Reese [2007(b)] employed the extended equations proposed by


Brenner (BNSE) [2005(a)], to predict the shock structure of Argon gas at various
Mach numbers. When they tried to solve the shock structure problem with the BNSE,
they could not obtain convergence. They had reasoned that since the continuity
equation proposed by Brenner [2005(a)], shown in equation (2.1), did not have any
diffusion term, the shock wave simulations could not lead to a converged solution and
subsequently, they had to replace the mass velocity by the volume velocity, or in other
terms, include diffusion terms in the continuity equation in order to obtain converged
solutions. In Figure 7.1, the shock structures obtained by the CNSE and the modified
Brenner Navier-Stokes equations (MBNSE) have been compared with the
144
experimental measurements by Alsmeyer [1976]. The dimensionless distance through
the shock, X1 has been obtained by normalizing the axial distance along the shock by
the molecular mean free path at the upstream of the shock l M 1 . Further, the density
and temperature profiles have been non-dimensionalised as shown in equation (7.18).
As observed in this figure, the MBNSE predicted the density profiles accurately in the
downstream of the shock. However, the prediction was poor in the upstream direction
and both the calculated temperature and density profiles were sharper in comparison to
the experimental measurements, as shown in the encircled region in Figure 7.1.

Further, Greenshields and Reese [2007(b)] also compared a number of other shock
structure parameters calculated using the MBNSE. They concluded that though the
numerical predictions obtained with the MBNSE were superior to the CNSE, the
prediction in the upstream region of the shock was poor. Since the ENSE proposed in
this thesis take into account the diffusion transport of mass accurately, it was felt that
prediction of shock structures would be a good test case to prove the correctness of the
equations. The results obtained with the ENSE were compared with the profiles
obtained with experimental measurements and other numerical techniques available in
published literature and one can observe that the results obtained with the ENSE are
much superior to that predicted with the MBNSE.

7.2 Computational inconsistencies faced in other methods

0.008 l M 1

0.083 lM 1

Cell sizes : 0.083 l M 1 , 0.056 l M 1 , 0.033 l M 1 , 0.021 l M 1 , 0.014 l M 1 , 0.008 l M 1

Dimensionless distance through the shock

Figure 7.2 Mach number, U1 T profiles obtained with various cell sizes by
Greenshields and Reese [2007(b)]; Volume diffusion coefficient was assumed to be
equal to thermal diffusivity
145
Before explaining the numerical solution procedure and the shock structures obtained
with the ENSE, it is necessary to review the problems faced by Greenshields and Reese
[2007(a), (b)] in predicting hypersonic shocks employing the MBNSE. As mentioned
earlier in the previous section, by employing the BNSE, as shown in equations (2.1),
(3.81) and (3.83)–(3.85), Greenshields and Reese [2007(b)] could not obtain converged
solutions of shock waves. It was necessary for them to incorporate the volume velocity,
as defined by equation (3.80), in the continuity equation instead of the originally
proposed mass velocity to obtain the converged solution.

Though the convergence issue was solved based on the above mentioned correction,
Greenshields and Reese [2007(b)], subsequently, faced with another problem when they
applied the diffusion coefficient to be equal to the thermal diffusivity of the gas, as
suggested by Brenner [2005(a)] and shown in equation (3.80). As the grid size was
reduced, an overshoot developed at the upstream edge of the profiles of all flow
variables and this overshoot continued to grow as the grid size became smaller. In
Figure 7.2, one such overshoot in the profile of Mach number, observed by
Greenshields and Reese [2007(a), (b)], is shown. In this figure, the axial distance along
the shock has been non-dimensionalised using the molecular mean free path at the
upstream of the shock l M 1 . They attributed the presence of this overshoot to the
additional term in the momentum flux, but not to the additional term in the energy flux
in the MBNSE and postulated that the overshoot might be caused by inappropriately-
calculated large volume diffusivity.

Alternately, they tried to replace the thermal diffusivity by the self-diffusivity


coefficient Dm of a monoatomic non-polar gas derived from Chapman-Enskog theory

using the Lennard-Jones potential which turned out to be Dm  1.32  1.32 . The

unphysical overshoot continued to remain even after this modification and finally
when the volume diffusivity was made equal to the kinematic viscosity of the fluid,

i.e. Dm  , the overshoot disappeared. The stability analysis carried out by

Greenshields and Reese [2007(b)] also showed that the solution was stable only when
the diffusion coefficient is close to the kinematic viscosity. It is interesting to note
that the arbitrarily obtained value of the diffusion coefficient in this analysis by
Greenshields and Reese [2007(b)] is the same as the one obtained by the systematic
derivations of the ENSE proposed in this thesis. As it can be observed in equation
(3.7), the self-diffusion coefficient was taken to be the kinematic viscosity of the fluid.

Interestingly, even after the above mentioned major changes to the BNSE,
Greenshields and Reese [2007(b)] could not obtain smooth profiles of flow variables
in the upstream leg of the shock structure as shown in Figure 7.1. With increasing
Mach number, the discrepancy of the computed density and temperature profiles with
the experimental measurements in the upstream direction increased continuously.
When one compares the diffusion transport proposed in the present work and by
146
Brenner [2005,(a)], as shown in equations (3.9) and (3.80), respectively, it is evident
that only the diffusion due to Fick’s term was considered by Brenner [2005(a)] with a
different diffusion coefficient but the Soret term, given by equation (3.11), had been
omitted completely. In the present derivations, both the Fick’s and Soret diffusion
terms have been included in the self-diffusion transport of mass and more importantly,
this has been properly incorporated in the continuity equation, as shown in equation
(3.38). Therefore, the ENSE proposed in this thesis, obtained by physically sound
derivations, are expected to predict the shock structures accurately and the results
obtained from these simulations are discussed below in this section.

7.3 Geometry and boundary conditions

In Figure 7.3, the geometry and boundary conditions employed in the simulations of
shock waves using the ENSE are shown. The the ENSE, derived in chapter 3, are given
below once again for the sake of clarity.

Inlet Symmetry
Pin Outlet
Po
m in Symmetry
T0 L  35 l M1

Figure 7.3 Geometry and boundary conditions employed in the shock wave simulations

Continuity equation:
  U iC  m iD
  (3.38)
t xi xi

Momentum equation:

 U Cj   U C
i U Cj    P   T
ij
 g j (3.40)
t x i x j x i

Total energy equation:


 E   U iC E 

q iT
  Ek
T
 C
m iD  PU iC   ij U j
 
 
 g jU Cj
 (3.64)
t xi xi xi xi xi

Total molecular momentum transport:


 U Cj U iC  2 C
  
T
ij     ij  U k  m iDU Cj  m DjU iC  2  ij m kDU kC (3.37)
 xi x j  3 xk 3
 

147
Total molecular transport of heat:
T
q iT    m iD e (3.21)
x i

The governing equations mentioned above can be simplified as given below omitting
the transient and gravity-related terms not relevant to the problem under consideration.

Continuity equation:
U iC  m D
 i (7.1)
xi xi

Momentum equation:
 U iCU Cj     U Cj U iC  2 C 
P
         ij  U k  m iDU Cj  m DjU iC  2  ij m kDU kC 
x i x j xi   xi x j  3
 x k 3 

(7.2)
Total energy equation:
 U iC E  m iD  PU iC    ij U j 
T C
  T D 
     m i C v T   E k   (7.3)
xi xi  xi  xi xi xi

Since the employed computational fluid dynamics solver, Fluent 12.0.16 has been
developed to solve only the CNSE, the additional terms in the ENSE were treated as
volumetric source terms and incorporated into the solution procedure through User-
Defined-Functions (UDF). It is important to note that Greenshields and Reese
[2007(b)] obtained solutions to shock structure problem using the nondimensional form
of the governing equations and boundary conditions. However, the present results were
obtained with the dimensional form of the governing equations, mentioned in equations
(7.4) – (7.8) and the following boundary conditions:

(i) In the experiments of Alsmeyer [1976], the static pressure and temperature at
the inlet was specifed to be 6.665 Pa and 300 K, respectively. Since the exact
dimensional boundary conditions employed by Greenshields and Reese
[2007(b)] were not available, the inlet conditions of Alsmeyer [1976] were
assumed for all the simulations. Since the comparisons are made in terms of
dimensionless variables, the error due to this assumption is expected to be less.
(ii) The other boundary conditions can be calculated using the Rankine-Hugoniot
relationships as described here. Since the inlet Mach number was specified in
the simulations, the inlet velocity U inC could be calculated based on the
following formula:
U inC  Ma in  Tin (7.4)

148
In equation (7.4), Ma in is the specified Mach number at the inlet,  the ratio of
the specific heats of the gas,  the specific gas constant and Tin the static
temperature at the inlet.
(iii) Based on the calculated inlet velocity employing equation (7.4), the velocity in
C
the downstream of the shock U out , was calculated as:
C
U out 1 3 
 1   (7.5)
C
U in 4  Main2 
(iv) Further, the outlet pressure was calculated using the following formula:
Pout 1
Pin 4

 5Main2  1  (7.6)

(v) Furthermore, the static temperature ratio across the shock was expressed as:
C
Tout Pout U out
 (7.7)
Tin Pin U inC
(vi) The molecular mean free path at the upstream of the shock l M 1 was calculated,
as shown in Greenshields and Reese [2007(b)]:
16  in Cin
lM 1  (7.8)
5 2 Pin
In equation (7.8),  is the specific heat ratio of the gas and  in and Cin are the
dynamic viscosity and velocity of sound at the inlet, respectively. One could also
calculate the molecular mean free path at the upstream of the shock l M 1 as given
below:
Tin
lM 1  (7.9)
2 d m2 N A Pin
In equation (7.9),  is the gas constant and d m and N A are the atomic diameter
and the Avagadro number, respectively. For the case of Argon, the atomic
diameter d m was taken to be 3.77  10 10 m . The molecular mean free path values
calculated by using equations (7.8) and (7.9) differed only by 1%.
(vii) Based on the specified velocity and density at the inlet, the mass flow rate at the
inlet was calculated and the same was specified in the simulations, as shown in
Figure 7.3.

As shown in Greenshields and Reese [2007(b)], the total length of the shock wave
geometry was taken to be 33 l M 1 . They observed that the solution obtained with a grid
count of 800 was found to be within 1% of the solution extrapolated to an infinitely
small mesh size. Therefore, in the present study, no grid-dependency study was carried
out and the grid count was chosen to be 1000 with the grid size of 0.033 l M1 . Second
order discretization schemes for the flow variables and the density-based-solver (DBS)
available in Fluent 12.0.16 were employed in the simulations.

149
7.3.1 Variation of dynamic viscosity with temperature

Greenshields and Reese [2007(b)] carried out a detailed analysis on the variation of
dynamic viscosity with temperature because selection of the right viscosity model is
crucial in shock wave simulations carried out to validate any hydrodynamic model.
Since the variation of viscosity with respect to temperature has a very strong influence
on the profile of a simulated shock, a wrongly chosen viscosity model can mask the
results obtained from the hydrodynamic model. Greenshields and Reese [2007(b)]
chose the following power law model of dynamic viscosity variation with temperature:
  AT S (7.10)
where S can be calculated as:
1 2
S  (7.11)
2 v 1

In equation (7.10) and (7.11), A is the constant of proportionality and v, the power law
coefficient in the intermolecular collision model. There are two theoretical limiting
cases for the intermolecular force law, (i) ν=∞, s=1/2 corresponds to hard-sphere
molecules and (ii) ν=5, s=1 represents the so-called Maxwellian molecules. However,
real molecules generally have been found to have value of ν in the range of 5 – 15. It
can be observed from equation (7.9) that increasing the value of the exponent S in a
shock structure simulation will introduce more momentum and energy diffusion,
particularly at the high-temperature downstream side of the shock. Needless to
mention, this additional diffusion will tend to smooth out the shock layer by increasing
its thickness in the downstream region. Therefore, as mentioned earlier, it is important
to set the correct value for the exponent S in equation (7.10) in order to investigate the
validity of any hydrodynamic model.

As discussed in Greenshields and Reese [2007(b)], Maitland and Smith [1972]


critically assessed the viscosities of a number of gases obtained from several sources.
For monoatomic gases such as argon, helium and xenon, they developed a power law
model for the viscosity with an estimated accuracy of 1.5% in the temperature range
80–2000 K; the upper limit of 2000K is the downstream temperature of a shock of
Mach 4.3 propagating into a room temperature gas. Further, Amdur and Mason
[1958] estimated the intermolecular potential at higher temperatures from observations
of the scattering of high-velocity molecular beams and produced tables of the viscosity
of gases at temperatures up to 15,000 K, corresponding to a shock of Mach 12.5. In
the more limited temperature range of 3500–8500 K, Aeschliman and Cambel [1970]
obtained the viscosity of argon with an accuracy of 12%. Similarly, correlations
between DSMC simulations and experimental shock density profiles were also
employed to obtain the exponent S, particularly at high temperatures; see Bird [1970],
Barcelo [1971], Steinhilper [1972], Sturtevant and Steinhilper [1974], Alsmeyer
[1976], Lumpkin and Chapman [1991] and Macrossan and Lilley [2003]. Based on
the above mentioned studies, one could estimate that the value of ν lies within a range
of 9  v  11corresponding to the power law exponent value S of 0.7  S  0.75 . The

150
most frequently employed value of S being 0.72, as quoted by Alsmeyer [1976].
Greensheilds and Reese [2007(b)] considered two temperature ranges to model the
viscosity of Argon in their study, (i) Mach numbers up to Mach 4.3 and (ii) Mach
numbers between 4.3 - 12.5. They figured out that the power law exponent, S could
be 0.68 and 0.76 in the two cases, respectively. Further, they also used the mean value
of 0.72 in all the simulations, as the control value.

Based on the assessment of various power law exponent values quoted in the
literature, it can be concluded that there is a certain uncertainty involved in the
prediction of dynamic viscosity variation with respect to temperature and this can be
as high as  10% in certain cases. It is necessary to consider this uncertainty while
comparing numerical results with the corresponding experimental measurements. In
all the simulations presented in this section, the power law exponent S was uniformly
chosen to be 0.72. Further, it is well-known that the Prandtl number of a mono-atomic
gas is 0.667 and hence, the local thermal conductivity  values in the simulations
were calculated from the estimated dynamic viscosity values, as given below.
C P
 (7.12)
Pr
where C P is the specific heat value of the gas at constant pressure.

7.4 Results and discussions

The ENSE, shown in equations (7.1)–(7.3), were employed to simulate shock


structures of argon gas with the following inlet Mach numbers, 2.31, 2.84, 3.8, 8, 9
and 11, respectively. The values of the Mach numbers were chosen to match the
experimental measurements of Alsmeyer [1976] and simulations of Greenshields and
Reese [2007(b)]. As discussed in section 7.2.1, Greenshields and Reese [2007(b)]
faced convergence difficulties because of (i) the absence of diffusion term in the
continuity equation of Brenner [2005(a), (b)] and (ii) the coefficient of the volume
diffusivity proposed by Brenner [2005(a)]. In the present study, there was
m D
convergence difficulty when the second term, Ek i in the right hand side of
xi
equation (7.3) was included in the energy equation. When this particular term was
omitted from the energy equation, the convergence of the simulation was excellent.
Since no reason could be identified for this behaviour, the simulations were carried out
without this particular term.

In order to compare the shock structures obtained in the present study with the
experimental measurements and other simulated profiles presented by Greenshields
and Reese [2007(a)], the following non-dimensionalisation was employed for all the
flow variables.
  in
  (7.13)
 out  in

151
By the above mentioned non-dimensionalisation, one could obtain a value of 0 and 1
at the upstream and downstream sections of the shock for all the flow variables,
respectively.

7.4.1 Density and temperature profiles

In Figure 7.4, the shock structures computed employing the CNSE and the ENSE are
shown for the inlet Mach number of 2.31. The experimental measurements of density
by Alsmeyer [1976] are also shown in this figure for comparison. It can be observed
from this figure that the shock structure predicted by the CNSE is sharper and the
profiles obtained with the ENSE are more diffuse and closer to the experimental
measurements. The difference in the shock structures obtained with the CNSE and the
ENSE is more in the downstream direction than the upstream region of the shock.

0.9
Nondimensional Density/Temperature

0.8

0.7

0.6 Rho_CNSE

0.5 Rho_ENSE

Rho_Expt
0.4
Temp_CNSE
0.3
Temp_ENSE
0.2

0.1

0
-7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7
Nondimensional distance

Figure 7.4 Comparison of simulated shock profiles obtained with the CNSE and the
ENSE and experimental measurements of Alsmeyer [1976]; Mach number = 2.31

It is well-known that in a shock structure, the density rises from its upstream value to
its downstream value behind the temperature and this is believed to be due to finite
relaxation times for momentum and energy transport, see Greenshields and Reese
[2007(a) and (b)]. However, there is no experimental verification of this phenomenon
since measurement of temperature profiles in a shock wave is nearly impossible. The
distance between the density and temperature profiles is known as the temperature–
density separation  TP . It can be visually observed from Figure 7.4 that the
temperature–density separation obtained by the CNSE is almost uniform across the
152
shock. On the contrary, the ENSE predicted more temperature–density separation in
the upstream as compared to the downstream of the shock.
1

0.9
Nondimensional Density/Temperature

0.8

0.7
Rho_CNSE
0.6
Rho_ENSE
0.5 Rho_MBNSE
0.4 Rho_Expt
Temp_CNSE
0.3
Temp_ENSE
0.2 Temp_MBNSE

0.1

0
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6

Nondimensional distance

Figure 7.5 Comparison of simulated shock profiles with the CNSE, the ENSE and the
MBNSE and experimental measurements of Torecki and Walenta [1993]; Inlet Mach
number = 2.84

1
Nondimensional Density/Temperature

0.9

0.8

0.7

0.6 Rho_CNSE
Rho_ENSE
0.5
Rho_Expt
0.4
Temp_CNSE
0.3
Temp_ENSE
0.2

0.1

0
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5
Nondimensional distance

Figure 7.6 Comparison of simulated shock profiles with the CNSE and the ENSE and
the experimental measurements of Alsmeyer [1976]; Inlet Mach number = 3.8

153
Similarly, the shock structures obtained with the CNSE and the ENSE are compared in
Figure 7.5 for the inlet Mach number of 2.84. The profiles obtained by Greenshields
and Reese [2007(b)] employing the MBNSE and the experimental measurements of
Torecki and Walenta [1993] are also shown in this figure for comparison. One can
observe from Figure 7.5 that the dimensionless density profiles obtained with the
ENSE and the MBNSE are closer to the experimental measurements whereas the
CNSE predict a much sharper profile. Similar to Figure 7.4, the difference between
the profiles obtained with the CNSE and the ENSE is more in the downstream of the
shock than the upstream region. Further, the CNSE and the MBNSE predicted a
higher temperature–density separation in comparison to the ENSE proposed in this
thesis. Moreover, the separation distance obtained with the ENSE is also not uniform
across the shock, as also discussed in Figure 7.4.

In Figure 7.6, the computed shock structures obtained with the CNSE and the ENSE are
compared with the experimental measurements of Alsmeyer [1976]. It can be observed
in this figure that the dimensionless density profile obtained by the ENSE is almost
close to the experimental measurements with a good spread whereas the profile
predicted by the CNSE is much sharper and differs from the experimental
measurements in both the upstream and downstream of the shock. It can also be
observed in Figure 7.6 that the nondimensional temperature profile obtained by the
ENSE also shows a very good spread in the upstream direction of the shock which is
absent in the profile calculated employing the CNSE.

The simulated dimensionless density and temperature profiles of the shock with the inlet
Mach number of 8 and 9 are shown in Figures 7.7 and 7.8, respectively. The profiles
obtained with the CNSE and the ENSE are compared with the experimental
measurements of Steinhilper [1972]. Further, the simulated profiles obtained with the
MBNSE are also shown in these figures for comparison. The experimental data and the
profiles due to Brenner [2005] were extracted from Figures 8 and 9 of Greenshields and
Reese [2007(a)]. It can easily be concluded from these figures that as expected, the
CNSE do not predict the shock structures accurately.

It is interesting to note that though the MBNSE predicted the downstream region of the
shock well, the prediction of both the density and temperature profiles in the upstream
direction was very poor. The encircled region in Figure 7.7 is shown once again in
Figure 7.9 and it can be observed in this figure that the density and temperature profiles
obtained by the MBNSE were abruptly sharp in the upstream direction, to the extent
that the profiles can be concluded as unphysical. Greenshields and Reese [2007(a)] also
discussed this poor prediction of the shock structures in the upstream direction and
concluded the following.

154
1

0.9
Nondimensional Density/Temperature
0.8

0.7 Rho_CNSE

0.6 Rho_ENSE
Rho_MBNSE
0.5
Rho_Expt

0.4 Temp_CNSE
Temp_ENSE
0.3
Temp_MBNSE
0.2

0.1

0
-8 -6 -4 -2 0 2 4 6
Nondimensional distance

Figure 7.7 Comparison of simulated shock profiles with the CNSE, the ENSE and the
MBNSE and experimental measurements of Alsmeyer [1976]; Inlet Mach number = 8

0.9
Nondimensional Density/Temperature

0.8

0.7

0.6 Rho_CNSE
Rho_ENSE
0.5
Rho_MBNSE
0.4 Rho_Expt
Temp_CNSE
0.3 Temp_ENSE
Temp_MBNSE
0.2

0.1

0
-8 -6 -4 -2 0 2 4 6
Nondimensional distance

Figure 7.8 Comparison of simulated shock profiles with the CNSE, the ENSE and the
MBNSE and experimental measurements of Alsmeyer [1976]; Inlet Mach number = 9
155
They pointed out that the molecular velocity distribution function in the upstream
region of the shock would be a non-Maxwellian combination of slow cold upstream
molecules and fast hot molecules that had diffused from downstream regions. They
argued with certainty that the hydrodynamic models such as the MBNSE would not be
able to treat this non-Maxwellian distribution of molecules properly and hence the
predictions of density and temperature profiles in this region were poor. Further,
Greenshields and Reese [2007(a)] also stated that a major advantage of the DSMC
technique in simulating intermediate-Knudsen number flows was its ability to produce
non-Maxwellian velocity distributions. It was further mentioned that apart from the
DSMC simulations, one could also predict the upstream region of the shock accurately
by employing certain extended hydrodynamic models of OKn 2 , such as the BGK
model and the Burnett equations etc., see Lumpkin and Chapman [1991], Reese
[1993] and Reese et al [1995]. It is well known that employing these complex
hydrodynamic models would involve huge amount of computational cost and there are
known problems related to instabilities. Further, the numerical implementation is also
difficult due to the large number of additional nonlinear and higher order derivatives.

0.2

0.18 Rho_CNSE
Nondimensional Density/Temperature

Rho_ENSE
0.16
Rho_MBNSE
0.14 Rho_Expt

0.12 Temp_CNSE
Temp_ENSE
0.1
Temp_MBNSE
0.08

0.06

0.04

0.02

0
-8 -7 -6 -5 -4 -3 -2 -1
Nondimensional distance

Figure 7.9 Comparison of shock structures within the encircled region shown in
Figure 7.7

However, it is interesting to observe that the density and temperature profiles obtained
with the ENSE predict the upstream region of the shock exceedingly well and the
agreement between the density profiles obtained with the ENSE and the experimental
measurements is excellent, as shown in Figures (7.7) – (7.9). This excellent
agreement between the predictions based on the continuum based the ENSE and the
156
experimental measurements raises serious questions about the existing assumptions
and reasons for the poor predictions of the prevalent hydrodynamic models. Later in
this section, the physics of diffusion driven mass transport in the shock wave is
explained in detail and one can notice there that there is no assumption regarding non-
Maxwellian distribution of molecular velocities and the assumption of local
equilibrium condition has also been maintained in the present analysis. In spite of
these assumptions, the ENSE were able to predict most of the salient characteristics of
normal shock waves accurately.

0.9
Nondimensional Density/Temperature

0.8

0.7

0.6 Rho_CNSE

Rho_ENSE
0.5
Rho_MBNSE
0.4
Rho_DSMC
0.3 Temp_CNSE

0.2 Temp_ENSE

Temp_DSMC
0.1
Temp_MBNSE
0
-10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5
Nondimensional distance

Figure 7.10 Comparison of simulated shock profiles with the CNSE, the ENSE and the
MBNSE and the DSMC simulations; Inlet Mach number = 11

In Figure 7.10, the shock structures of a normal shock with the inlet Mach number of 11
are shown based on various computational models. The results obtained with the
CNSE, the ENSE and the MBNSE are shown for comparison along with the DSMC
simulations of Lumpkin and Chapman [1991]. It can be observed that the prediction of
density profile by the MBNSE in the upstream region of the shock is poor, similar to the
unphysical sharp profiles shown in Figure 7.8. The deviation of density profile
predicted by the ENSE from that of the DSMC simulations is also found to be
significant in the upstream region of the shock. However, the density profile obtained
in the present study is quite smooth and close to the profile obtained by the DSMC
simulations in the far upstream region of the shock.

It is obvious to note the significant deviation in the temperature profiles obtained with
the ENSE and the DSMC simulations in Figure 7.10. The distance between the
157
temperature and density profiles obtained with the ENSE is almost half of that predicted
by the DSMC simulations. It is to be noted here that unlike the density profiles, it is not
easy to experimentally measure the temperature profiles in strong shock waves.
Therefore, one is forced to accept the temperature profiles obtained by simulation
techniques such as the DSMC. It is well-known that there are a number of modelling
assumptions employed in these simulations and one may have doubts on the accuracy of
these results.

7.4.2 Velocity and Mach number profiles

The dimensionless convective velocity and normalized Mach number profiles obtained
with the CNSE and the ENSE are compared in Figures 7.11 and 7.12. One can observe
in Figure 7.11 that the spread of the convective velocity profile obtained with the ENSE
is more than that predicted with the CNSE. In order to understand the reasons for this
difference in velocity profiles, one needs to consider the two mechanisms of mass
transport involved in the shock wave simulations employing the ENSE. It is well-
known that the convective mass transport takes place from the upstream direction
towards the downstream of the shock wave. Interestingly, the diffusive transport of
mass is in the opposite direction to the convective transport. Let us analyse the
diffusion transport of mass in detail since it is the most important characteristics in
shock waves which is revealed by the simulations carried out employing the ENSE.

From the definition of diffusion transport of mass, shown in equation (3.9), one can find
out that the diffusion mass transport occurs in the opposite direction to the density and
temperature gradients and in shock waves, it transpires to be from the downstream of
the shock towards the upstream direction. Therefore, the convective and diffusion mass
transports oppose each other in shock waves and the diffusion mass transport causes the
hot downstream fluid to move towards the colder upstream region which increases the
spread of temperature and density profiles. In Figure 7.12, the convective and diffusion
velocities are plotted across the shock for different inlet Mach numbers. It is evident
from this figure that the diffusion velocity is present only in the region of the shock and
its value is always negative. The maximum value of the diffusion velocity is around
100m/s and 380m/s for the inlet Mach number values of 2.34 and 8, respectively.
Needless to mention, such a strong diffusion mass transport will influence the
characteristics of the shock structures considerably, as evidently shown in Figures 7.4 –
7.10.

In Figure 7.12, the ratio of diffusion to convective mass transports is also plotted for the
case of two inlet Mach numbers and it can be observed that the maximum value of this
ratio is around 20%. The ratio of mass flow rates smoothly varies across the shock and
reaches the value of zero both in the upstream and downstream regions. Since the
values of temperature and density are constants in both the upstream and downstream
directions, the gradients tend to zero and hence the diffusion transport of mass is
negligible away from the shock.

158
1

0.9
CNSE
Nondimensional velocity

0.8 ENSE
0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5
Nondimensional distance
(a)

0.9
CNSE
0.8 ENSE
Nondimensional velocity

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
-5 -4 -3 -2 -1 0 1 2 3 4 5
Nondimensional distance
(b)

Figure 7.11 Comparison of dimensionless velocity profiles obtained with the CNSE and
the ENSE; Inlet Mach number (a) 2.84 and (b) 8

159
1000 0.18

900 Diffusion_velocity
0.16

Ratio of diffusion to convection mass flow


800 Convection_velocity
Convection / Diffusion velocity, m/s

Diffusion_Convection_Ratio 0.14
700

600 0.12

500
0.1
400
0.08
300

200 0.06

100
0.04
0
0.02
-100

-200 0
-10 -8 -6 -4 -2 0 2 4 6 8 10
Nondimensional distance
(a)
3000 0.25
Diffusion_velocity

2500 Convection_velocity

Ratio of diffusion to convection mass flow


Convection / Diffusion velocity, m/s

Diffusion_Convection_Ratio 0.2

2000

0.15
1500

1000
0.1

500

0.05
0

-500 0
-10 -8 -6 -4 -2 0 2 4 6 8 10
Nondimensional distance
(b)

Figure 7.12 Convection and diffusion velocity profiles obtained with the ENSE; Inlet
Mach number (a) 2.84 and (b) 8

160
Further, the ratio of diffusion to convective mass transports is also shown for all the
Mach numbers studied in this chapter in Figure 7.13. It can be observed in this figure
that the maximum value of the ratio is around 12.5% for the Mach number of 2.31 and
the value rises to around 21.5% for the Mach number of 11. Though the ratio increases
with increasing Mach number, it stabilizes around 21-22% for larger Mach numbers.
Further, the maximum value of the ratio does not occur exactly at the middle of the
shock but in a region slightly upstream to it.

0.25
Ratio of diffusion to convection mass transport

Ma
2.31
0.2
2.84
3.8
8
0.15
9
11

0.1

0.05

0
-10 -8 -6 -4 -2 0 2 4 6
Dimensionless distance across the shock

Figure 7.13 Ratio of diffusion to convective mass transports in shock waves obtained
with the ENSE

As discussed earlier, Greenshields and Reese [2007(a)] speculated that the inability of
the MBNSE to incorporate the non-Maxwellian behaviour of molecules was the reason
for the poor prediction of density and temperature profiles at higher values of inlet
Mach numbers, as shown in Figure 7.7 – 7.10. As reasoned by them, the non-
Maxwellian behaviour of the molecules was supposed to be caused by the mixing of
slow-moving cold upstream molecules with faster hot molecules trickling from the
downstream direction. This was stated to be the reason for the non-equilibrium
condition prevailing in shock waves. In general, one may fathom the movement of hot
molecules of the downstream region towards the upstream direction because of random
motion of atoms/molecules. However, one can notice from Figures 7.12 and 7.13 that
the diffusion velocity, moving towards upstream from the downstream direction, is
predictable, non-random and governed by physically sound constitutive relationship.
Therefore, one might argue that achieving local equilibrium condition is possible in

161
shock waves in spite of the mixing of molecules from hot and cold regions and it is
hoped that this line of reasoning with the help of the ENSE will grasp the interest of
researchers.

1 0.25

0.9
CNSE
0.8 0.2
Normalised Mach number

ENSE_conv._vel

Diffusion Mach number


0.7 ENSE_total_vel
ENSE_diff_vel
0.6 0.15

0.5

0.4 0.1

0.3

0.2 0.05

0.1

0 0
-10 -8 -6 -4 -2 0 2 4 6 8 10
Nondimensional distance
(a)
1 0.35

0.9 CNSE
0.3
ENSE_conv_vel
Normalised Mach number

0.8

Diffusion Mach number


ENSE_total_vel
0.7 0.25
ENSE_diff_vel
0.6
0.2
0.5
0.15
0.4

0.3 0.1
0.2
0.05
0.1

0 0
-12 -10 -8 -6 -4 -2 0 2 4
Nondimensional distance
(b)

Figure 7.14 Comparison of normalized Mach number profiles obtained with the CNSE
and the ENSE; Inlet Mach number (a) 2.84 and (b) 8

In Figure 7.14, the normalized Mach number profiles obtained with the CNSE and the
ENSE are shown at two different inlet Mach numbers. For the case of the ENSE, both
162
the convective and total velocity values were normalized by the local velocity of sound
to obtain the convective and total Mach numbers, respectively. As observed in this
figure, the Mach number profile obtained by the CNSE is very sharp whereas the spread
of the Mach number profiles obtained with the ENSE is very high. Further, the
deviation of the predictions based on ENSE from the classical solutions increases with
increasing Mach number. Since the ENSE positively influence the spread of the
temperature profiles, the Mach number profile also depicts similar behaviour.

Further, it can also be observed that the difference between the normalized convective
and total Mach number profiles is small. It needs to be stressed here that this is caused
by the scaling employed in the normalization of the flow variables, as shown in
equation (7.18). Actually, the diffusion Mach number is significant and as shown in
Figure 7.14, the maximum value of diffusion Mach number is 0.21 and 0.33 for the
cases of inlet Mach number values of 2.84 and 8, respectively. Therefore, the difference
between the total and convective Mach numbers is also significant but this has been
masked by the normalization employed in the analysis. Further, it can also be noticed
that the downstream region of the shock profile is not completely visible in the
normalized Mach number profiles, i.e., the region of the shock where the dimensionless
distance > 0 and this region has also been masked by the normalization.

7.4.3 Inverse density thickness, L1

Apart from the comparisons of shock structures, one can compare the results obtained
from different methods based on other shock parameters. As the first parameter, the
dimensionless shock inverse density thickness was chosen which is defined as given
below.
lM 1
L1   max
(7.14)
 2  1
where l M 1 is the molecular mean free path in the upstream of the shock region and 1
and  2 are the densities at the upstream and downstream regions, respectively. Further,
 max represents the maximum value of the density gradient within the shock. In the
case of shock wave problem, there is no easily determinable characteristic length scale
and hence, it is not easy to define the Knudsen number. It is customary to use the
actual thickness of the shock layer itself as the characteristic length scale. Therefore,
the dimensional inverse density thickness can be used to define the Knudsen number
of the shock wave, as
l
Kn   max (7.15)
 2  1
where l is the local molecular mean free path of the shock wave. Greenshields and
Reese [2007(a)] observed that based on the above mentioned definition of the Knudsen
number Kn , one would obtain very high Knudsen numbers and hence any
hydrodynamic model should fail to predict the characteristics of the shock structures.

163
They, however, reasoned that one could still assess any extended hydrodynamic
equations for their usefulness as engineering models.

0.5

0.45
Inverse density thickness, 1/m

0.4

0.35

0.3

0.25

0.2

CNSE
0.15
ENSE
Experiment
0.1
MBNSE_S=0.72
0.05
1 2 3 4 5 6 7 8 9 10 11
Mach number

Figure 7.15 Comparisons of inverse density thickness, L1 of the shock structure
obtained with different methods

In Figure 7.15, the dimensionless inverse density thickness L1 values obtained with
different models are compared. The values obtained with the CNSE and the ENSE are
shown along with the values predicted by the MBNSE. Greenshields and Reese
[2007(a)] employed the MBNSE and predicted the dimensionless inverse density
thickness L1 values with three values of the power law exponent, S of dynamic
viscosity of Argon. In Figure 7.15, the values obtained by the power law exponent S of
0.72 are only shown. Further, Alsmeyer [1976] provided the most comprehensive
collection of experimental shock data, comprising previously published work as well
as his own results. The dimensionless inverse density thickness L1 values collated
from the experimental measurements from Steinhilper [1972], Alsmeyer [1976] and
Torecki and Walenta [1993] are shown in Figure 7.15, as shown in Figure 10 of
Greenshields and Reese [2007(a)].

As it can be observed in this figure, the CNSE fail to predict this important parameter
of the shock wave characteristics and the values predicted are much more than the
experimental predictions. The values obtained with the MBNSE with the power law
exponent of 0.72 closely match the experimental measurements. It can be observed in
this figure that the ENSE predict 8-10% higher values of inverse density thickness as
164
compared to the experimental measurements. Though it may be considered that the
agreement of the numerical results obtained with the ENSE and experimental
measurements is reasonable, it is intuitively felt that one can improve the predictions
of the ENSE further because of the following reasons.

As observed by Greenshields and Reese [2007(a)], the numerical predictions of the


inverse density thickness was found to be influenced by the chosen power law
exponent since with increasing value of the power law exponent, the prediction was
found to improve and resemble the experimental measurements. They also found out
that the power law exponent S of 0.74 seemed to represent the best-fit line of the
experimental measurements of the inverse density thickness, shown in Figure 7.15. It
can also be observed in this figure that the predicted values of inverse density
thickness by the ENSE are closer to the experimental measurements at smaller inlet
Mach numbers than the larger values. Since the chosen power law exponent S of the
viscosity in this study was 0.72 and this value was found to agree well with the
experimental measurements of viscosity at lower temperatures. In the simulations by
Greenshields and Reese [2007(a)], they mentioned that the power law exponent S of
0.76 was found to be suitable for inlet Mach number values more than 8. In short, if
one chooses the right viscosity model, the predictions of the ENSE can be improved
further. Since the objective of the studies mentioned in this chapter is to show that the
ENSE can be employed to predict the characteristics of shock waves accurately and
the self-diffusion of mass is the sole reason for the enhanced spread of shock profiles
observed in the experiments. Therefore, the additional studies involving choosing the
correct power law model of viscosity was not carried out, as part of this thesis.

Since the temperature in the upstream region of the shock is very low, it is not correct
to choose a single power law exponent S for the whole shock wave. It is required to
use different values of the power law exponent at different sections of the shock wave
depending on the temperature range prevailing in the region. Further, there are other
viscosity models, such as Sutherland’s model and it may be possible that the
predictions can be improved further by employing these models, as well. Moreover,
there are some other pertinent issues with the ENSE as discussed later in this chapter,
which demand some more detailed attention. The author is interested to work on each
of the above mentioned aspects in order to improve the predictions of shock waves by
employing the ENSE.

7.4.4 Density asymmetry quotient, Q 

Another important quality parameter of shock wave predictions is known as the density
asymmetry quotient Q  which is defined as,
0

  dx

Q  
(7.16)
 1   dx

165
where   is the local value of dimensionless density. Since the inverse density
thickness comparison, shown in Figure 7.15, depends only on the density gradient at the
midpoint of the shock structure, it does not express anything about the overall shape of
the profile. Therefore, the density asymmetry quotient Q  can be used to describe the
complete shock profile and some experimental data is also available for this parameter
for effective comparisons.
1.5

1.4
Density Asymmetry Quotient

CNSE
1.3 ENSE
Experiment
MBNSE_1
1.2 MBNSE_2

1.1

0.9
1 2 3 4 5 6 7 8 9 10 11
Mach number

Figure 7.16 Comparisons of density asymmetry quotient Q  of the shock structure


obtained with different methods

As shown in equation (7.16), the density asymmetry quotient Q  is a measure of the


skewness of dimensionless density profile of shocks about the midpoint. Therefore, a
symmetric shock would consequently have the density asymmetry quotient Q  of unity
but real shock waves are found to be not completely symmetrical about their midpoint.
It is observed from the experimental measurements that the density profile is skewed a
little towards the downstream direction at smaller inlet Mach numbers, i.e. Q  1 .
However, as the Mach number increases, the spread of the shock increases, as
discussed earlier. Hence, the value of the density asymmetry coefficient Q  becomes
more than unity and it increases with increasing inlet Mach number within the range
studied in the experiments.

In Figure 7.16, the density asymmetry quotient Q  values obtained with the CNSE and
the ENSE are compared with the experimental measurements of Alsmeyer [1976].
Further, the values obtained by the MBNSE, the data extracted from Figures 11 and 4
166
from Greenshields and Reese [2007(a)] and [2007(b)], respectively, are also shown in
the same figure. Since the plots of density asymmetry quotient employing the
MBNSE differed significantly in Greenshields and Reese [2007(a)] and [2007(b)],
both profiles are presented in Figure 7.16 as the MBNSE_1 and the MBNSE_2,
respectively. It is evidently seen in Figure 7.16 that both the CNSE and the MBNSE
do not at all predict the asymmetry in the density profile properly. The CNSE over-
predict the density asymmetry quotient Q by ~30% whereas MBNSE_1 under-
predict it by ~20%. The profile shown as the MBNSE_2 over-predicts the density
asymmetry quotient at lower Mach numbers and it approaches the experimental
measurements at higher inlet Mach numbers. However, the predictions by the ENSE
agree excellently with the experimental measurements at all Mach numbers. This
result indicates that the ENSE developed in this thesis can be one of the stronger
competitors for predicting the structure of the shock waves accurately.

7.4.5 Temperature-density separation,  T

3.5
Dimensionless temperature - density seperation distance

CNSE
3 ENSE
DSMC
Brenner
2.5

1.5

0.5
1 2 3 4 5 6 7 8 9 10 11
Mach number

Figure 7.17 Comparisons of dimensionless temperature - density separation distance


 T of the shock structure obtained with different methods

Another shock structure parameter is known as the temperature–density separation  T


and this distance is measured between the midpoints of the normalized density and
temperature profiles. It is understood from the available literature that the shock
structure of density happens to be in the downstream of that of temperature and this is
believed to be due to the finite relaxation times for momentum and energy transport.
However, there is no experimental data for this parameter since it is very difficult to
167
measure temperature profiles in a shock. Typically, one has to employ the DSMC
simulations to generate this data for comparison with hydrodynamic models.

In Figure 7.17, the calculated temperature–density separation distances obtained by the


CNSE, the ENSE and the MBNSE are compared against the values obtained by the
DSMC simulations of Lumpkin and Chapman [1991]. It is interesting to note that
none of the models are able to predict the temperature–density separation distance
calculated by the DSMC simulations. The separation thickness  T calculated by the
DSMC simulations increases monotonically with Mach number. The values calculated
from all the hydrodynamic models are lower than those predicted by the DSMC
simulations for almost all the Mach numbers. The predictions of the CNSE and the
MBNSE show a decreasing trend up to the Mach number of ~2.3 and then the values
gradually increase with increasing Mach number. The separation thickness values
calculated from the ENSE also increase monotonically with Mach number similar to
the DSMC results, i.e. both the profiles are almost parallel, but the absolute values are
much less in comparison to the DSMC simulations.

Though the differences between the results obtained by the DSMC simulations and the
ENSE are strikingly evident, one needs to consider the following before concluding
about the suitability of the ENSE in predicting shock structures. In the comparisons
with experimental measurements of inverse density thickness L1 and density
asymmetry quotient Q  , the ENSE performed excellently, as shown in Figures 7.15 and
7.16. The temperature–density separation thickness is not an experimentally
measurable parameter and it might be possible that the underlying assumptions in the
DSMC simulations could be the reason for this large discrepancy observed in the
comparisons shown in Figure 7.17. At this juncture, it is difficult to favour one method
over the other based on the set of comparisons provided in this chapter. However, the
availability of the ENSE can revive the studies on shock structures and characteristics
and it is believed that promising new understanding of the underlying physics of shock
waves can be obtained from such studies.

7.4.6 Discussions on the definition of Knudsen number

In the absence of solid walls, it is always difficult to choose the correct characteristic
length to define dimensionless parameters and the definition of Knudsen number for the
case of shock waves is no exception. It is well-known for a long time that the CNSE do
not predict shock structures accurately and historically, the high Knudsen number of the
flow was blamed as the reason for the failure of the CNSE. It was argued that the local
equilibrium assumption would no longer be valid in high Knudsen number flows and
hence the continuum based CNSE was expected to fail in the predictions. Based on this
argument, many complex extensions were proposed to the CNSE to incorporate the
experimentally observed spread of the shock structures in the computational
simulations. In order to understand the characteristics of the shock structures and

168
choose the right model to simulate the shock waves, it was mandatory to select a
suitable characteristic length to define the Knudsen number.

Knudsen number, defined by equation (7.15) 0.35


Ma 2.31
0.3 Ma 2.84
Ma 3.8
Ma 8
0.25 Ma 9
Ma 11
0.2

0.15

0.1

0.05

0
-10 -8 -6 -4 -2 0 2 4 6
Nondimensional distance
(a)
Knudsen number, defined by equation (7.17)

0.25
Ma 2.31
Ma 2.84
0.2 Ma 3.8
Ma 8
Ma 9
Ma 11
0.15

0.1

0.05

0
-10 -8 -6 -4 -2 0 2 4 6
Nondimensional distance
(b)

Figure 7.18 Comparison of Knudsen number profiles at different inlet Mach numbers

As mentioned by Greenshields and Reese [2007(a)], it is customary to choose the


Knudsen number definition given by equation (7.15) in shock wave simulations. The
Knudsen number profiles, calculated using equation (7.15), are shown in Figure 7.18(a)
for various inlet Mach numbers. It can be observed in this figure that the Knudsen
number profiles resemble the profiles of dimensional variables shown in Figures (7.4)–
(7.10). For a given Mach number, the maximum value of the Knudsen number is at the
169
upstream section of the shock where the pressure is the minimum. Though the lower
temperature prevailing at the upstream section of the shock attempts to reduce the
Knudsen number, the influence of pressure is stronger. The minimum Knudsen number
occurs at the downstream region of the shock, as expected. However, it is possible to
observe the following ambiguity in the profiles shown in Figure 7.18(a).

The inlet pressure and temperature were the same for all the Mach numbers simulated in
Figure 7.18(a) and therefore, the molecular mean free path computed using equation
(7.14) is also identical for all the cases. Since the selected characteristic length, based
on the maximum value of normalized density gradient, is a function of the inlet Mach
number, one obtains different Knudsen number values at the upstream of the shock. It
was felt that this does not characterize the shock wave correctly. Therefore, the
Knudsen number was redefined using the local diffusion length scale and molecular
mean free path l, as given by,
 1  1 T 
Kn  l    (7.17)
  x1 2T x1 

The calculated Knudsen number profiles employing equation (7.17) are plotted at
different inlet Mach number values in Figure 7.18(b). It is evident from this figure that
the Knudsen number values identically tend to zero in both upstream and downstream
regions for all the Mach numbers. As the inlet Mach number increases, one might
expect that the tendency to deviate from local equilibrium condition and the peak value
of Knudsen number profile will also increase. However, as observed in Figure 7.18(b),
such a monotonic increase in the peak Knudsen number is not observed and the
maximum value of Knudsen number is limited to ~0.22 only at all inlet Mach numbers.
It is argued that this particular characteristic of the shock wave indicates the reason why
the excellent prediction of shock structures was possible by employing the continuum-
based the ENSE proposed in this thesis.

7.4.7 Discussions on violation of laws of thermodynamics

It is, in general, believed that as the flow becomes rarefied, the predictions of
continuum formulations, such as the CNSE, become inaccurate since these
formulations employ linear approximations to the molecular momentum and heat
transport, as shown in equations (2.24) and (2.30), respectively. Therefore, higher
order terms in the constitutive relationships of molecular momentum and heat
transports were incorporated in the governing equations to improve predictive
capabilities under rarefied conditions and so the second-order systems of
hydrodynamic equations, such as the Burnett and the Woods equations, were obtained.
While it was observed that these equations provided a better description of the shock
structure on coarse grids, they were prone to instabilities, believed to be small
wavelength instabilities, when the grids were refined. It was conjectured that an
inherent entropy inconsistency might be the cause of computational instability and
subsequently it was proven that the one-dimensional Burnett equations, when applied
170
to the hypersonic shock structure problem, can violate the second law of
thermodynamics as the local Knudsen number increases above a critical limit. For a
detailed description of the above mentioned problem, one can refer to Balakrishnan
[2004].

S 
Dimensionless specific entropy,

Dimensionless length, x1 l M 1 Dimensionless length, x1 l M 1

Figure 7.19 Dimensionless specific entropy, S  profiles of a shock with inlet Mach
number of 20, as obtained by Balakrishnan [2004]; Grid fineness ratios, x1 l M 1 :
(a) 8 and (b) 4

In Figure 7.19, S indicates the change of entropy. As shown in Figure 7.19, the
dimensionless entropy S  profiles portrayed a dip in the upstream region of the
shock and the extent of this dip increased with increasing grid fineness ratio x l M 1 .
For an inlet Mach number of 20, it was found that the critical grid fineness ratio x l M 1
is 4 and the temperature became negative for values less than 4. In Figure 7.20, the
dimensionless entropy profiles of a shock wave, calculated with the CNSE and the
ENSE, with an inlet Mach number of 11 are shown. Interestingly, the grid fineness
ratio x l M 1 employed in the simulations was 0.033 and no negative entropy dip was
observed in the upstream region of the shock. Subsequently, the grid fineness ratio
x l M 1 was decreased to a value of 0.01 and even then, no abnormalities in the specific
entropy profiles were observed. Further, the entropy profiles obtained with a number
of inlet Mach numbers are also presented in Figure 7.21 and the solutions obtained
were entropy-consistent in each and every case. As explained by Balakrishnan [2004],
by employing the modified Burnett equations, one could choose the right set of
equations and coefficients to obtain an entropy consistent shock structure. Even after
that elaborate careful selection of terms, one could go up to a grid fineness ratio x l M 1
of only 0.1 and the solution showed inconsistency below the value of 0.1. On the

171
contrary, the ENSE based solutions faced no such constraints and the solution was
entropy-consistent at all grid fineness levels.
5

4.5
Dimensionless specific entropy

3.5

2.5 CNSE
ENSE
2

1.5

0.5

0
-16 -14 -12 -10 -8 -6 -4 -2 0 2 4
Nondimensional distance

Figure 7.20 Dimensionless specific entropy, S  profiles obtained with the CNSE
and the ENSE of a shock with inlet Mach number of 11; Grid fineness ratio
x l M 1  0.033

4.5
Ma 2.31
4
Ma 2.84
Dimensionless specific entropy

3.5 Ma 3.8
Ma 8
3
Ma 9
2.5 Ma 11

1.5

0.5

0
-16 -14 -12 -10 -8 -6 -4 -2 0 2 4 6 8
Nondimensional distance

Figure 7.21 Dimensionless specific entropy, S  profiles of a shock obtained with the
ENSE with different inlet Mach numbers; Grid fineness ratio x l M 1  0.033
172
1140

1120

1100
Total temperature, K

CNSE
1080 ENSE

1060

1040

1020

1000
-10 -8 -6 -4 -2 0 2 4 6 8 10
Nondimensional distance
(a)
6900

6800

6700

6600
Total temperature, K

CNSE
6500 ENSE
6400

6300

6200

6100

6000

5900
-10 -8 -6 -4 -2 0 2 4 6 8 10
Nondimensional distance
(b)

Figure 7.22 Comparison of total temperature profiles obtained with the CNSE and the
ENSE; Inlet Mach number (a) 2.34 and (b) 8

One more interesting observation about the shock wave results can be made related to
the first law of thermodynamics. Shock waves can, in general, be considered as
adiabatic systems and one dimensional shock waves are certainly adiabatic. It is well-
known that the total temperature at all sections must remain constant in adiabatic flows.
In the case of the CNSE, the total temperature of the fluid T0 is defined by the following
equation:
173
T0 T 
U 
C 2
1
(7.18)
2C P

In equation (7.18), T and U 1C are the local static temperature and convective velocity of
the fluid, respectively. Similarly, the total temperature of the fluid T0 can be defined for
the case of the ENSE as,

T0 T 
U 
T 2
1
(7.19)
2C P
where U 1T is the local total velocity of the fluid. The total velocity is the vector sum of
the convective and diffusion velocities as shown in equation (3.51). In Figure (7.22),
the total temperature profiles obtained with the CNSE and the ENSE are presented for
two inlet Mach numbers. It is interesting to note that the total temperature does not
remain constant and reaches a maximum value at the shock in the case of the CNSE.
One can note that it is an undisputable violation of the first law of thermodynamics and
the author does not know any published literature discussing this discrepancy.
Surprisingly, the total temperature does also not remain constant in the case of the
ENSE. In this case, the total temperature drops at the shock to a minimum value and
this is also a violation of the first law of thermodynamics.

It is known that the CNSE do not predict shock structures properly and hence it is no
wonder that the results obtained with the equations show certain thermodynamic
inconsistency. However, the predictions of the ENSE were pretty close to the
experimental measurements and therefore it is surprising to find similar thermodynamic
inconsistency in this case as well. There may be a number of reasons for this
inconsistency and few of the possibilities are enlisted here.
(i) The selection of the correct power law exponent of the viscosity model can
influence the spread of the shock wave and therefore, can reduce the deviation
from the thermodynamically consistent condition of constant total temperature
across the shock. Further, one need not assume constant power law exponent
throughout the shock for a given Mach number and can choose the exponent
based on the local temperature values, as discussed before.
(ii) In the simulations presented in this chapter, Argon was assumed to be adhering
to the ideal gas law. Since very high temperatures are encountered in the
downstream region of strong shock waves, one can expect some deviation from
the ideal gas law. One can incorporate real gas behaviour in the computations
in order to study its influence on this observed violation of the first law of
thermodynamics.
(iii) More importantly, as discussed in the beginning of section 7.2.3, the second
m D
term, Ek i in the right hand side of equation (7.3) had to be dropped from the
xi
total energy equation and only then converged solutoin could be obtained with
the ENSE. The reason for this behaviour of the equations is not known yet.
174
(iv) Further, the thermal energy equation, shown in equation (3.56), is given once
again here with a slight modification.


 
 e   U iC e

 
  
T D 
 m i e   P
U i T
  ij
U j
(7.20)
t x i x i  x i  x i x i
In the last two terms in the right hand side of equation (7.20), U i has been
introduced instead of the convective velocity U iC . Since the diffusion velocity
is a macroscopic quantity like the convective velocity, it can also take part in the
expansion work and viscous dissipation. Therefore, in equation (7.20), U i can
be replaced with U iT in one or both of the terms. The author tried to change the
two terms in terms of the total velocity to see whether the consistency with
respect to the first law of thermodynamics could be achieved but faced with
poor quality of predictions of the shock structures.

Based on the above mentioned discussions, it can be concluded that the problem of
numerical computation of shock structures is not yet understood completely and there
remains a number of open questions. However, based on the results discussed in this
chapter, it needs to be mentioned here that the ENSE can be employed as a new and
effective tool to understand this complex flow structure.

175
176
Chapter 8

IMPORTANT RESULTS AND SUGGESTIONS


FOR FURTHER RESEARCH

8.1 Introductory remarks

The governing equations of fluid mechanics, known as the CNSE were, in general,
considered to be well-developed and capable of handling all sorts of complex problems
accurately. The excellent agreement between the experimental measurements and
theoretical and computational predictions were deemed to be the proof of exactness of
the governing equations and the employed constitutive relationships of diffusion
transports of momentum and heat, provided by the Newton’s law of viscosity and
Fourier’s law of heat diffusion, respectively. Interestingly, certain problems, such as
gas flows through microchannels and capillaries, accurate predictions of shock
structures, physically convincing explanations to thermophoresis and thermal
transpiration etc., remained unexplained by the CNSE. These problems were thought to
be in the fringe region where the continuum assumption, upon which the CNSE were
based, itself was not valid and hence special treatments, such as invoking ‘slip-flow’
theory in the case of microchannel flows and introduction of extended hydrodynamic
models such as Burnett equations in the case of shock wave simulations etc., were
provided to solve these so called ‘high Knudsen number flows’. Since one could
obtain excellent agreements with experimental measurements in most of the flow cases,
the completeness of the CNSE was never questioned in the past.

In the recent past, there is an increasing discomfort amongst the scientific community
about the comprehensiveness of the CNSE and there were number recent attempts to
derive what is known as ‘the extended Navier-Stokes equations’ in order to broaden the
validity of the continuum based the CNSE, the most recent one was proposed by
Brenner [2005(a)]. Though there was a general consensus on ‘something is amiss’ with
the CNSE, no missing physical process was identified as the reason for the observed
deviations of theoretical predictions from corresponding experimental measurements
and hence, the proposals were extemporized and remained fundamentally questionable,
for example the ‘volume diffusion’ theory proposed by Brenner [2005(a)]. As claimed
by Brenner [2005(a)], it was possible for the volume to diffuse without any
corresponding mass diffusion and this theory had been accepted because of the
explanation provided to problems such as thermophoresis, even though the basic
assumption was flawed. At the same time Brenner [2005(a)] proposed his version of
the extended equations (BNSE), Durst et al [2006] proposed the ENSE based on the
argument that the self-diffusion transport of mass needed to be incorporated in the
continuity equation. Though the presence of self-diffusion in liquids and gases was
known for centuries, it was not included in the CNSE because of its apparent smallness.
Durst et al [2006] argued that though the self-diffusion transport of mass remained
negligible in comparison to the convective transport in most flow situations, it could
assume greater significance in gas flows where strong density and temperature gradients
were present such as gas flows through microchannels and shock waves. Further, the
additional diffusive transports of momentum and energy caused by the self-diffusion
mass transport were also incorporated in the momentum and energy equations,
respectively and subsequently, the ENSE were obtained. In the present thesis, the
ENSE were derived based on the procedure followed by Durst et al [2006], as shown in
chapter 3. Further, comparisons between the ENSE proposed in this thesis and the
BNSE were also provided and the absence of diffusive transport in the continuity
equation was stated to be the major flow in the derivations by Brenner [2005(a)]. Any
theoretical model needs to be validated with experimental measurements in a number of
flow situations to prove its validity. Therefore, the ENSE proposed in this thesis was
also employed to solve problems such as gas flows through straight microchannels and
capillaries, gas flows through microchannels with a backward facing step and
predictions of shock structures in a monoatomic gas.

In chapter 4, the characteristics of gas flows through straight microchannels were


numerically computed and compared with the experimental measurements of Arkilic et
al [2001] and Maurer et al [2003]. The analytical solutions obtained with the CNSE
and the no-slip boundary condition at the solid wall proved that the convective
acceleration terms could be neglected in gas flows through straight microchannels. The
numerical predictions obtained with the ENSE agreed excellently with the experimental
measurements of Maurer et al [2003]. This conclusively proved that the reason for the
observed additional mass transport in the experiments was indeed the self-diffusion
mass transport driven by the density gradient in the channel and not because of the
widely believed and accepted Maxwell-slip based ‘slip-flow’ theory. It could be
observed in this chapter that both the convective and diffusion velocity profiles obtained
with the ENSE satisfied the no-slip boundary condition at the wall. However, since the
diffusion velocity reached its maximum value immediately at the outer region of the
wall roughness elements, the total velocity profile, the vector sum of the convective and
diffusion velocities, resembled the slip-flow profile. Further, it was stressed that it was
essential to understand that the origin to this slip-like behaviour observed in the
simulations with the ENSE was different from the ‘slip-flow’ theory. In the case of the
‘slip-flow’ theory, it was necessary to conduct experiments to measure the value of slip-
coefficient whereas the diffusion velocity was completely deterministic from the first
principles. Based on the ENSE, it was possible to predict the ‘Knudsen paradox’,
defined as the occurrence of the minimum in the conductance profile, and physical
reasons behind the apparent ‘paradoxical’ behaviour of the conductance profile were
also explained in detail. Since there was an excellent agreement between numerical
predictions based on the ENSE and experimental measurements, it could be argued that
gas flows through microchannels, in the range studied in this thesis, were essentially
within the continuum region. Further, to elaborate this aspect, the Knudsen number was
defined based on the characteristic length obtained from the ‘diffusion length scale’ of
the flow instead of the often employed height of the channel. Interestingly, it was
178
observed that the Knudsen number defined based on the diffusion length scale was two
to three orders of magnitude smaller than that obtained based on the classical definition
as shown in Figure 4.13.

Subsequently, based on the insights obtained from the numerical simulations of


microchannel flows employing the ENSE, analytical solutions to gas flows through
microchannels and capillaries were attempted, as described in chapter 5. The
convective acceleration terms were dropped from the momentum equations based on the
detailed order of magnitude analysis. The resulting equation could be integrated to
obtain the analytical solutions for the mass flow rate and velocity through
microchannels and capillaries. More significantly, a characteristic pressure PC was
introduced which was found to describe the nature of flow characteristics through
microchannels and capillaries accurately. When the local dimensionless pressure
obtained based on the characteristic pressure, P PC was equal to unity, the convective
and diffusive mass transports were found to be equal. Similarly, when the local
pressure was more than the characteristic pressure PC , then the convective mass
transport was dominant over the diffusion transport and vice versa. The mass flow rates
obtained by the analytical solutions for both microchannels and capillaries were found
to agree excellently with the corresponding experimental measurements. Furthermore,
one semi-analytical solution procedure was developed to obtain the pressure profiles
along the length of the channel and capillaries and the theoretically obtained profiles
agreed remarkably with the experimental values, as well. It can be convincingly stated
here that one can use the analytical solutions of the ENSE to obtain all the
characteristics of isothermal gas flows through microchannels and capillaries accurately,
up to an outlet Knudsen number of unity.

One of the major bottlenecks of the ‘slip-flow’ theory is its inability to predict gas flows
through complex microchannel geometries, say for example, a backward facing step.
Since the tangential momentum accommodation coefficient needs to be determined
experimentally in order to employ the ‘slip-flow’ theory, one would find it difficult to
determine the same in complex geometries. This inability of the ‘slip-flow’ theory led
to the application of the Lattice Boltzmann Simulations (LBS) and the Direct
Simulation Monte Carlo (DSMC) simulations in complex microchannel geometries.
Interestingly, the ENSE based simulations could predict gas flow characteristics through
microchannels with a backward facing step exceedingly well and the excellent
comparisons were obtained with the DSMC simulations over a wide Knudsen number
range. It was found that the velocity values at the wall reported in the DSMC
simulations were not evaluated exactly ‘at the wall’ but a small normal distance away
from it. This led to some discrepancy in comparing the wall velocity magnitudes
obtained with the ENSE. Other than this minor discrepancy, all the flow characteristics
of microchannel flows with a backward facing step were accurately determined
employing the ENSE.

179
As shown in the derivations of the ENSE in this thesis, the self-diffusion mass flow rate
does also depend on the temperature gradients present in gas flows and hence, it was
necessary to assess the equations in predicting gas flows with strong temperature
gradients. Prediction of one dimensional shock structures of supersonic and hypersonic
gas flows of a monoatomic gas was taken to be the test problem for which extensive
experimental measurements and results from the DSMC simulations are available for
comparison. Further, the numerical predictions, based on the modified extended
Navier-Stokes equations proposed by Brenner [2005(a)] (MBNSE), provided by
Greenshields and Reese [2007(a)], were also employed for comparisons. As explained
in chapter 7, the shock structures were predicted accurately by the ENSE for all the
Mach numbers studied in both upstream and downstream regions of the shock whereas
the MBNSE failed to predict the upstream characteristics of the shock accurately at
higher Mach numbers. Shock parameters obtained by the ENSE such as inverse density
thickness L1 and density asymmetry quotient Q  were found to be in excellent
agreement with the experimental measurements. However, the ENSE were found to
under-predict the temperature–density separation thickness  T in comparison to the
predictions of the DSMC simulations. Since this is not an experimentally determined
parameter, it is necessary to employ elaborate validations to verify the predictions of
temperature–density separation thickness  T based on both the ENSE and the DSMC
simulations. Furthermore, as detailed by Balakrishnan [2004], the shock structures
obtained employing the Burnett equations were found to violate the second law of
thermodynamics at the upstream location of the shock for lower grid fineness values.
The predictions by the ENSE were free of this thermodynamic violation at any grid
fineness values. More importantly, the simulations carried out employing both the
CNSE and the ENSE were found to violate the first law of thermodynamics as the total
temperature was found to vary along the length of the shock. One can notice that the
characteristics of shock structures have not yet been understood completely and it is
stressed here that the ENSE can be employed as a useful tool in exploring the
underlying physics of hypersonic shock waves.

8.2 Outlook towards future research

It is felt that the problems solved in this thesis provide a strong foundation for analyzing
many diffusion dominated gas flows in detail. In this section, some of the potential
research areas that can be explored employing the ENSE are outlined.

8.2.1 Gas flows through microchannels

One can observe from the various results discussed in chapters 4 and 5 that the problem
of isothermal gas flows through straight microchannels and capillaries has been
completely resolved employing the ENSE. It has also been completely proven that the
additional mass flow rates observed in experiments is indeed the self-diffusion transport
of mass and not because of the Maxwellian slip at the solid boundary. It is felt that the
following problems can be dealt further in the case of straight microchannels.
180
(i) Gas flows through straight channels with heat transfer: A number of
experimental measurements are available in the published literautre in the area
of heat transfer in microchannel gas flows. The ENSE can be employed to
predict the heat transfer rates in microchannels and compare with the
experimental measurements.
(ii) It is generally felt that the noncontinuum effects will be very significant even at
the Knudsen number value of around 0.01-0.05. On the contrary, the ENSE,
which was derived based on the continuum hypothesis, were found to work
satisfactorily at Knudsen numbers of O(1) , as shown in chapters 4 and 5. One
can compare experimental measurements at higher Knudsen numbers with the
simulated results to determine the upper limit of the validity of the ENSE.
(iii) The ENSE were found to satisfactorily predict the characteristics of gas flows in
microchannels with a single backward facing step. However, if the
experimental measurements are made available, one can evaluate gas flows in
very complex flow situations. As of today, there is no other continuum based
models available which can handle such complex flows efficiently and further
employing the DSMC simulations or any other molecule-based simulations will
be computationally prohibitive.

8.2.2 Computations of hypersonic shock structures

Needless to mention, the shock structures obtained with the ENSE were superior to the
MBNSE at all values of inlet Mach number. The shock structures, obtained with the
ENSE, in the upstream direction of the shock wave was found to be smooth and closer
to the experimental measurements whereas the profiles simulated with the MBNSE
could not calculate the upstream region properly. However, it was felt that the
predictions of density profiles could be improved further if the correct self-diffusion
coefficient in the simulations with the ENSE. As mentioned by Greenshields and Reese
[2007(a)], the self-diffusion coefficient was equal to Dm  1.32  1.32   when
calculated using the Chapman–Enskog theory employing the Lennard – Jones potential.
However, they faced convergence difficulties while using this value of the diffusion
coefficient and finally, settled for Dm      .

When one employs the classical kinetic theory of gases, the self-diffusion coefficient is
predicted to be equal to the kinematic viscosity of the gas, i.e., Dm   as employed in
the simulations presented in chapters 4–7. However, the real atoms and molecules do
not behave exactly identical to the assumptions of the kinetic theory of gases and the
self-diffusion coefficient is modelled, under these conditions, as Dm  C where C
varies between 1.2 – 1.5; see Kennard [1938]. As stated by Kennard [1938], the value
of the constant C was found to be 1.2 and 1.543 when the hard spheres assumption and
the inter-molecular repulsion based on the 5th power of the distance were considered,
respectively.

181
1

Nondimensional Density/Temperature 0.9

0.8

0.7
Rho_CNSE
0.6
Rho_ENSE
0.5
Rho_MBNSE
0.4 Rho_Expt
0.3 Temp_CNSE

0.2 Temp_ENSE

0.1 Temp_MBNSE

0
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5
Nondimensional distance
(a)
1

0.9
Nondimensional Density/Temperature

0.8

0.7

0.6
Rho_CNSE
0.5 Rho_ENSE
Rho_MBNSE
0.4
Rho_DSMC
0.3 Temp_CNSE
Temp_ENSE
0.2
Temp_DSMC
0.1 Temp_MBNSE

0
-10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5
Nondimensional distance
(b)

Figure 8.1 Comparison of shock structures obtained with modified self-diffusion


coefficient; Inlet Mach number (a) 2.84 and (b) 11

In order to verify the influence of the modified self-diffusion coefficient, predictions of


shock structures were done for two Mach numbers with the self-diffusion coefficient of
Dm  1.32 , mentioned by Greenshields and Reese[2007(a)] and the modified self-
diffusion mass transport was calculated as,
182
 1  1 T 
m iD  1.32    (8.1)
  xi 2T xi 

The shock structures calculated using the modified self-diffusion coefficient are
presented in Figures 8.1. While comparing with the shock structures presented in
Figures 7.5 and 7.10, it can be noticed that the agreement between the density profiles
obtained with the ENSE and experimental measurements is excellent in Figure 8.1(a).
Similarly, the profiles obtained with the ENSE also compare favourably with the values
obtained by the DSMC simulations in Figure 8.1(b). This interesting result clearly
indicates that the right selection of the self-diffusion coefficient can indeed assist in
predicting shock structures accurately employing the ENSE. This result again confirms
that the self-diffusion mass transport is the solitary reason for the observed spread of
shock structures in experimental measurements and one need to consider the ENSE in
order to obtain physically consistent shock structures in hypersonic flows.

Furthermore, it can be noted in chapter 4 that the height of microchannel was adjusted
slightly in the numerical simulations employing the ENSE to obtain exact agreements
with the experimental measurements. If the modified self-diffusion coefficient had been
employed, such a correction in the channel dimension would not have been required and
similar to shock wave simulations; one would obtain excellent agreement with the
experiments in simulations of microchannels as well.

Apart from the above mentioned domain of research in hypersonic shock waves
employing the ENSE, the following areas can also be explored further.

(i) As indicated in Figure 7.22, the total temperature profiles obtained by both the
ENSE and the CNSE do not remain constant across the shock wave and violate
the first law of thermodynamics. It is intuitively argued that by choosing the
correct form of the energy equation, as indicated in equation (7.24), one can
address this violation positively.
m D
(ii) Further, the term Ek i was neglected in the total energy equation presented
xi
in equation (7.8). When this particular term was included, it was not possible to
obtain converged shock structures employing the ENSE and the exact reasons
for this behaviour has not yet been understood. It is necessary to revisit the
derivations of the total energy equation and also the solution procedure to find
out the underlying reasons for this problem.
(iii) Interestingly, the temperature – density sepearation distance  T predicted by
the ENSE did not agree with the values obtained by the DSMC simulations
whereas there was excellent agreement of the inverse density thickness L1 and
density asymmetry quotient Q obtained by the ENSE and experimental
measurements. It is widely accepted in the compressible flow literature that the
density lacks behind temperature in a hypersonic shock, though there are no
183
experimental verification of this phenomenon. This particular problem is
considered to be an excellent puzzle related to fundamentals of compressible
flow physics which needs to be solved.
(iv) Further, the viloation of the second law of thermodynamics in the case of the
Burnett’s equations was discussed in the upstream region of shock waves, as
shown in Figure 7.19. As stated in section 7.4.7, it was demonstrated that one
could easily mitigate this thermodynamic violation by employing the ENSE.
Interestingly, one can also observe another entropy inconsistency in the
downstream region of the shock in Figure 7.19. It can be observed in this figure
that the entropy reaches a maximum before reaching a slightly lesser value in
the downstream of the shock. Further, it can be noticed in Figure 7.20 that this
hump is less in the case of the ENSE than that of the CNSE. It is obvious that
the entropy can only increase or stay constant across the shock and this
observed hump is generally overseen in the literature. When the upstream
entropy inconsistency was discussed elaborately by Balakrishnan [2004], there
was no mention of this downstream violation. However, it is clearly evident
that this is another violation of second law of thermodynamics and it is
intuitively stated here that this violation can also be removed by choosing the
correct form of the energy equation.

8.2.3 Thermophoresis

Thermophoresis is an observed physical phenomenon and needs to be considered while


studying deposition of soot particles in industrial gas cleaning applications,
determination of exhaust gas particle trajectories from combustion devices, study of
particulate material deposition on turbine blades, studies on the efficiency of air filters
and aerosol scrubbers and also in a number of other areas involving deposition of
particles such as chemical vapour deposition, see Davis [2006], Garg and Jayaraj
[1988], Homsy et al [1981]. Thermophoresis is defined as the motion of non-Brownian
particles, suspended in an otherwise quiescent medium, from hotter to colder regions,
i.e. the particles move against the externally imposed temperature gradient on the static
fluid. Brenner [2005(a)] proposed his extended Navier-Stokes equations only to explain
this phenomenon and he attributed thermophoretic particle motion, in the so-called
‘‘near-continuum’’ range of Knudsen numbers Kn  1 , to small non-continuum
effects. He further elaborated that these non-continuum effects are caused by the action
of thermal stresses existing in the gas proximate to the particle surface, resulting in the
Maxwell slip (‘thermal creep’) at the surface. It is well-known that the Maxwell slip
condition is usually applied to flow cases where the Knudsen number is large, i.e.
Kn>0.01. On the contrary, the phenomenon of thermophoresis can be observed in
quiescent gases at all pressures subjected to temperature gradients where the continuum
assumption is very much valid and the Knudsen number is very small, i.e. Kn <<1.

In spite of this obvious incongruity, as of today, thermophoretic movement of aerosol


particles in gases is accepted as a strictly non-continuum phenomenon since no such
motion, either of the particle or the fluid, in a quiescent stratified gas subjected to a
184
temperature gradient is predicted by the CNSE with the traditional no-slip boundary
condition at the particle surface. Recently, Brenner [2005(a)] developed a continuum
theory to explain thermophoretic motions observed in gases based on the introduction of
a fictitious volume velocity, defined by equation (3.84), in place of the classical
convective velocity in the momentum equation as shown in equation (3.88). Further, he
also argued that the no-slip boundary condition at solid surfaces should also be in terms
of the volume velocity and this condition would provide the ‘slip-velocity’ at the solid
surface in terms of the mass velocity. It is not evident, at least to the author, how one
can experience the volume motion without any corresponding mass motion.
Interestingly, one can explain the phenomenon of thermophoretic motion employing the
ENSE. In this section, only an approach to this problem is outlined in line with the
ENSE and the author has intentions to work on this problem subsequently.

+H Hot wall

Reflected mass
from the wall
Thermophoretic
particle
2H

Self-diffusion mass
transport

-H Cold wall

Figure 8.2 Explanation of thermophoretic motion based on the ENSE

Let us consider a mono-atomic gas entrained between two parallel walls maintained at
two different constant temperatures, the hot wall maintained at a constant temperature
of TH on the top at a position of +H and the cold wall of temperature TC at the bottom at
a position of -H. The two plates are separated by a distance 2H and the pressure is
constant throughout the domain. Let us also assume that the gravity is not present and
hence the gas is expected to be stagnant from the continuum view point, apart from
molecular fluctuations. Under these conditions, based on the classical fluid mechanics,
the gas will experience a steady state conduction heat transfer given by Fourier’s law
dT T  TC
q     H . Now, let us place a non-Brownian spherical particle of radius
dx 2H
a, density  P and thermal conductivity  P . It is evident based on the CNSE that the
particle will be stationary in the absence of any macroscopic fluid motion and gravity.

185
However, as stated above, the particle tends to move from the hot to cold region, i.e.
from the top towards bottom as shown in Figure 8.2 due to thermophoretic motion.

It is interesting to note from equation (3.9) that the self-diffusion transport is not zero as
given below.
 1 dP 1 dT 
m 1D       (8.2)
 P dx1 2T dx1 
Further, since the pressure is constant in the domain, equation (8.2) can be simplified as,
 dT
m 1D  (8.3)
2T dx1
It can be observed that the Soret self-diffusion mass transport term is non-zero for the
case considered in Figure 8.2 and interestingly, the direction of this mass transport,
given by equation (8.3), is towards the hot wall and it is counter-intuitive to the
observed thermophoretic motion as the thermophoretic velocity of the particle is
towards the cold wall. It is evident from equation (8.3) that the molecules continuously
diffuse and move from colder regions towards hotter region of the continuum because
of the action of self-diffusion mass transport, driven by the temperature gradient. One
can notice that if the Soret self-diffusion mass transport is present alone in the system
and then based on the continuity equation, given by equation (3.44), the self-diffusion
dm 1D
mass transport will automatically become zero, i.e.  0  m 1D  constant  0 based
dx1
on the impermeability condition at the wall.

Since the self-diffusion transport of mass is not equal to zero when a finite temperature
gradient is present in the system, then there must be a balancing mass transport
mechanism to counteract the self-diffusion. Because of the continued action of self-
diffusion mass transport, more and more molecules will be forced to move near the hot
wall from the colder regions of the domain. Therefore, the density of the gas near the
hot wall will tend to increase and since the temperature is fixed to be a constant, the
local pressure near the hot wall will also tend to rise. However, such an increase in
pressure has not yet been observed near the hot wall in thermophoretic motion studies.
Interestingly, as stated by Brenner [2005(a)], while studying thermal transpiration in
long capillaries, Reynolds [1879] observed a pressure difference between the two ends
of the capillary, with the pressure being highest at the hotter end, i.e. PH  PC and
similar pressure differences were observed by many other researchers; see Bielenberg
and Brenner [2004]. It is well-known that pressure differences in open flow systems,
such as boundary layer flows, jet flows etc., are not significant whereas the same can be
observed in closed-conduit flows such as pipes and channel flows. The pressure
difference caused by self-diffusion mass transport at the hot wall will be miniscule and
hence it might not have been noticed in thermophoresis studies. The difference in
pressure may be measured only by placing extremely sensitive measuring instruments at
both hot and cold walls and such an experiment is deemed to be a valid proof for the
description of thermophoretic motion given below.
186
As the pressure near the hot wall tends to increase, one can deduce that convection
currents will set in opposing the self-diffusion mass transport, as shown schematically
in Figure 8.2. Or simply, one may consider that the hot wall reflects back the self-
diffusion current due to the impermeability condition and thereby balancing the mass
flow caused by the self-diffusion transport of mass and satisfying the extended
continuity equation shown below.
dU 1C dm 1D dU 1T
  0 (8.4)
dx1 dx1 dx1

It can be observed that the amount of reflected mass flow rate from the hot wall will be
limited by the self-diffusion transport or in other words, both the reflected and self-
diffusion mass transports are equal in magnitude and opposite in direction. Hence no
macroscopic fluid motion is observed under the conditions depicted in Figure 8.2. It is
stressed here that despite the presence of two opposing streams of macroscopic mass
flow, the gas system will pose as an quiescent medium with no appreciable motion, as
observed in experiments. This ‘perceived’ quiescence of the gas, undergoing steady one
dimensional heat conduction, has puzzled researchers over the decades and is the source
of many contradicting empirical and adhoc theories of thermophoretic motion.

Though the two macroscopic mass streams balance each other, the values of
momentum of the two streams are not equal. Since the density at the hot wall is lower
than that of the cold wall, for a given mass flow rate, the velocity of the stream
reflected from the hot wall will be more than that moving towards it from the cold
wall. Therefore, the momentum, given by the product of the mass flow and velocity,
towards the cold wall is more than that towards the hot wall. Or in other words, if a
particle were to present in between the hot and cold walls, it will experience a net
force towards the cold wall because of the excess momentum from the hot wall and
hence, the particle will move from the hotter to colder regions of the medium. It is
stressed here that the apparent quiescent gaseous medium, subjected to an externally
imposed temperature gradient, is not ‘calm and quiet’ as expected but it undergoes
continuous volumetric expansion associated with macroscopic mass movements or in
other words, the work done by the applied temperature gradient on the quiescent
medium is not zero but is a finite positive value.

The above mentioned description of the thermophoretic motion is physically sound


and can also be determined from the first principles. It is planned to take this
argument forward and derive the relationships for the thermophoretic force and
velocity and it is important to state here that the initial results are encouraging.

8.2.4 Thermal transpiration

Another interesting phenomenon observed experimentally is known as thermal


transpiration. When a gas filled capillary of known initial pressure, whose ends are
perfectly sealed, is subjected to an externally imposed temperature gradient in the
187
absence of gravity, the only known molecular transport is the Fourier’s law of heat
diffusion based on the classical fluid mechanics. However, Reynolds [1879] observed
in his experiments that an eventual pressure difference developed within the capillary
from the initial uniform pressure and the pressure was more in the hotter region of the
capillary and he coined this phenomenon as the thermal transpiration. This thermo-
molecular pressure phenomenon, have been confirmed by many others employing a
host of different gases; see Knudsen [1910], Los and Ferusson [1952] and Annis
[1972]). Needless to mention, the CNSE failed to predict this phenomenon and hence,
the thermal transpiration was thought to be caused by the non-continuum effects
driven by the Maxwell slip of molecules along the solid surface. Recently, Bielenberg
and Brenner [2006] applied the BNSE by Brenner [2005(a)], to explain the thermal
transpiration phenomenon. As stated in the previous subsection, Brenner [2005(a)]
did not provide a valid physical reason for the existence of volume velocity without
the corresponding mass motion and the volume conservation equation, see Bielenberg
and Brenner [2006], instead of the mass conservation continuity equation, is highly
questionable.

Wall of capillary (Adiabatic wall)


Hot Cold
wall R wall
T  TH Axis T  TC

L
Wall boundary conditions
  1  1 T 
U TT     
   x1 2T x1 
U NT  0
Figure 8.3 Geometry and boundary conditions of thermal transpiration simulations

It is argued emphatically here that the underlying fundamental transport phenomenon


of the thermal transpiration is nothing but the self-diffusion of mass. To demonstrate
this, a sample simulation was carried out with the geometry and boundary conditions
shown in Figure 8.3. The radius R and length L of the capillary was taken to be 10 μ m
and 0.0005 m, respectively. Further, the hot and cold walls were maintained at 500
and 300 K, respectively. Argon was taken to be the working fluid of the capillary.
The tangential total velocity U TT was specified, based on the self-diffusion mass flow
rate, on all wall boundaries and the normal total velocityU TN was taken to be zero due
to impermeability condition. The initial pressure in the capillary was 10,000 Pa.

The simulated temperature and pressure rise values, obtained with the ENSE, are
shown in Figure 8.4. It can be observed in this figure that there is a positive pressure
188
rise in the capillary and the rise is more near the hot wall as compared to the cold wall.
The average value of pressure rise obtained at the hot wall is 210 Pa whereas the same
at the cold wall is 42.5 Pa. Needless to mention, the simulated pressure and
temperature profiles are analogous to the experimental observations of Reynolds
[1879] and Knudsen [1910]. Further, it is also possible to explain the physics behind
the thermal transpiration based on the ENSE. In Figure 8.5, the computed velocity
profile at the mid-section of the capillary is shown. It is evident from this profile that
the self-diffusion mass flow rate, driven by the imposed temperature gradient on the
capillary, moves along the horizontal wall of the capillary from the cold wall towards
the hot wall. When this diffusion mass flow reaches the hot wall, it is reflected back,
similar to the explanation given for the thermophoresis in the previous subsection,
towards the cold wall and this reflected mass flow moves near the axis of the capillary.
Further, it is interesting to note that the net flow rate at any section of the capillary is
zero and hence, one will presume that the gas is quiescent and does not undergo any
macroscopic movement. Obviously, this insight into the gas-filled sealed capillary
flow cannot be obtained without employing the ENSE.
500 225

480 Temperature
200
Pressure difference
460
175
440
Temperature, K

Pressure rise, Pa
150
420

400 125

380
100
360
75
340
50
320

300 25
0 0.2 0.4 0.6 0.8 1
Dimensionless distance along the capillary

Figure 8.4 Computed pressure rise and temperature profiles of an Argon-filled


capillary, TH  500 K , TC  300 K and Pin  10,000 Pa

Further, there is another interesting phenomenon observed as described by Landau and


Lifschitz [1966]. When two gas filled chambers, maintained at two different
temperatures at the same initial pressure, are connected by a very thin capillary of
diameter D. When the capillary diameter D is much larger than the molecular mean
free path, i.e. d  l , the pressure in both the gas chambers were observed to be one and
the same, irrespective of the temperatures maintained in gas chambers. However, when

189
the diameter D became less than the molecular mean free path l , i.e. d  l , there was an
impending pressure rise of the hot chamber over the cold chamber. Further, from the
observations, the following relationship was obtained between the pressures and
temperatures.
PC P
 H (8.5)
TC TH

0.9

0.8
Dimensionless radial distance

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
-0.15 -0.1 -0.05 0 0.05 0.1 0.15
Axial velocity, m/s

Figure 8.5 Computed velocity profile of an Argon-filled capillary, TH  500 K ,


TC  300 K and Pin  10,000 Pa

Obviously, the CNSE and other hydrodynamic models do not predict the pressure rise
and its relationship with the temperature of the two chambers. However, the analysis
based on the ENSE can arrive at the relationship given by equation (8.5) very easily. As
we had seen in the previous two subsections, the net self-diffusion mass transport in the
system considered is zero. Therefore, one can write, employing equation (3.9), that,
 1 dP 1 dT 
m 1D        0 
d ln P  d ln T
 
P 
 Constant
 (8.6)
 P dx1 2T dx1  dx1 dx1 T
It is evident that one can obtain the relationship obtained from the experiments directly
using equation (8.6). Based on the above mentioned arguments, it can be stated that
many unresolved problems which were thought to be in fringe areas of the validity of
continuum assumption of fluid mechanics and necessarily explained by the so called
‘non-continuum’ or ‘high Knudsen number’ effects, are explained by the introduction of
self-diffusion mass transport into the governing equations. It is really surprising that
this well-known diffusion mass transport was not considered or incorporated into the
190
governing equations before embarking on the explanation of the experimentally
observed phenomena based on the ‘non-continuum’ effects.

8.2.5 Strongly heated pipe flows

In all the examples described in this thesis so far, it was generally believed that the non-
continuum effects were responsible for the observed deviations from experimental
measurements and the Knudsen number was, usually large. As stated in section 1.1, it
is usually observed that the CNSE are valid and accurate for all cases where the
continuum assumption is very much valid and the Knudsen number, based on the
characteristic dimension of geometry, is very small. In many gas flow cases of
engineering applications, the pressure remains very close to the atmospheric pressure
and hence the CNSE has been found to predict the flow characteristics accurately.
However, it is possible for certain flow conditions to exist where one may need to
consider the ENSE even under flow conditions where the CNSE were considered to be
accurate.

Shehata and McEligot [1998] carried out experimental measurements of heat transfer in
strongly heated pipes subjected to ‘near-uniform’ wall heat flux. Since the high
Reynolds number air flow through the pipe was subjected to very strong heat fluxes, the
buoyancy production of turbulence was expected to be very high and special turbulence
models were developed based on the experimental measurements by modelling the
buoyancy production and dissipation of turbulence according to the measured heat
transfer characteristics. Since in the experimental measurements of Shehata and
McEligot [1998], the air flow was subjected to very strong temperature gradients, it was
possible that the self-diffusion transport of mass could play a significant role in
determining convective heat transfer characteristics close to the pipe wall. Hence, the
ENSE were employed to simulate one of the experimental cases of air flow through
heated pipe given by Shehata and McEligot [1998].

In Figure 8.6, the local Nusselt number profiles obtained by the ENSE and the CNSE
are compared with the experimentally determined values. It is evident from this figure
that the CNSE results employing the k   turbulence model without any ‘special
treatments’ to turbulence production and dissipation terms, did not obtain the correct
local Nusselt number profiles and it over-predicted the values in comparison to the
measurements. The measured local Nusselt number value at the dimensionless axial
distance of 22.8 was 10.38 whereas the CNSE predicted the Nusselt number to be
17.18, a difference of 60%. At the same dimensionless location, the Nusselt number
predicted by the ENSE was observed to be 12.2 with the observed difference of only
15%. Further, the predictions by the ENSE were found to agree well with the
experimental values throughout the length of the pipe. Needless to mention, it is
evident that by disregarding the presence of self-diffusion mass transport, the influence
of self-diffusion was lumped into the buoyancy induced production and dissipation of
turbulence by either modifying parameters in the turbulent model and/or incorporating
some more additional terms in the governing equations of turbulence. More
191
importantly, this can even lead to poor predictions when the ‘faulty’ turbulent model is
employed to solve other problems. Therefore, it is essential to employ the ENSE for
solving even ‘large scale’ engineering gas flow problems with large pressure/density
and temperature gradients.

40

CNSE
35
ENSE
Experiment
30
Local Nusselt number

25

20

15

10

5
0 5 10 15 20 25
Nondimensional axial distance

Figure 8.6 Comparison of local Nusselt number profiles computed employing the
CNSE and the ENSE with experimental measurements of Shehata and McEligot
[1998]

It is noted here that the excellent agreement between the ENSE and experimental
measurements, as shown in Figure 8.6, were obtained despite the employed the ENSE
were not properly derived for turbulent flows. There can be some more additional terms
when the Reynolds averaged extended Navier-Stokes stokes equations (RAENS) are
obtained for turbulent flows similar to the Reynolds averaged classical Navier-Stokes
equations, famously known as the RANS equations. Further, based on some initial
analysis, it has been found that the axial conductance of heat along the length of pipe
cannot be neglected and hence the constant heat flux boundary condition, employed by
Ezato et al [1999] and Shehata and McEligot [1998] is only an approximation and not
physically feasible. It is believed that the predictions of heat transfer characteristics by
the ENSE of the experimental measurements of Shehata and McEligot [1998] can be
greatly improved if the above mentioned changes are incorporated.

192
8.3 Concluding Remarks

Although the CNSE are widely employed to predict the characteristics of all types of
gas flow problems with a great degree of accuracy, a number of problems remain
where the classical equations fail to provide a satisfactory explanation. Greenshields
and Reese [2007(a)] provided a detailed (may not be comprehensive, as stated by
them) list of such problems where certain degree of deviations have been found
between theoretical/numerical predictions with experimental measurements and are
given below.

(i) The shock structure problem;


(ii) The nonlinearity of the thermal and momentum Knudsen layers (the region
O(λ) close to solid surfaces);
(iii) The ‘Knudsen paradox’ in Poiseuille flow, i.e. the minimum in the mass flow
rate at around Kn ≈ 1, as well as bimodality in the temperature profile;
(iv) Drag coefficients and heat transfer in: flow around a sphere, flow around a
cylinder and Couette flow;
(v) Variation of skin friction on cones in supersonic flow;
(vi) Base pressures on cone-cylinder configurations in supersonic flow as the
Knudsen layer extends far into the wake region;
(vii) Thermophoresis, i.e. the force on particles suspended in a rarefied gas between
two parallel plates of different temperature;
(viii) Dispersion and absorption of ultrasonic sound waves.

It is interesting that many of the above mentioned problems have been solved with
excellent agreements with experimental measurements by employing the ENSE. Based
on the elaborate discussions in the preceding chapters, it is claimed that the ENSE is the
set of equations that encompass all classes of flow problems and the CNSE are a subset
of the ENSE which can be derived by neglecting few of the terms from the ENSE.
Hence one may wonder whether it is apt to describe the CNSE as ‘the abridged set of
Navier-Stokes equations’ instead of identifying the equations proposed in this thesis as
‘the extended Navier-Stokes equations’.

193
194
REFERENCES

1. Aeschliman D. P. and Cambel A. B., ‚Properties of gases at very high


temperatures‘, Physics of Fluids, 13, p.2466, 1970.

2. Agrawal A., Djenidi L. and Antonia R. A., ‘Simulation of gas flow in


microchannels with a sudden expansion or contraction’, Journal of Fluid
Mechanics, vol.530, pp.135-144, 2005.

3. Agrawal A., Djenidi L. and Agrawal A., ‘Simulation of gas flow in microchannels
with a single 90obend’, Computers and Fluids,
doi:10.1016/j.compfluid.2009.01.004, 2009.

4. Alsmeyer H., ‘Density profiles in argon and nitrogen shock waves measured by the
absorption of an electron beam’, Journal of Fluid Mechanics, vol. 74, Issue 03,
pp.497-513, April 1976.

5. Amdur I. and Mason E. A., ‘Properties of gases at very high temperatures’,


Physics of Fluids, 1, p.370, 1958.

6. Arkilic E. B. and Schmidt M. A., ‘Gaseous slip flow in long microchannels’,


Journal of Micro-electro-mechanical systems, Vol.6, No.2, pp.167-178, June,
1997.

7. Arkilic E.B., Schmidt, M. and Breuer, K., ‘Gaseous flow in microchannels’,


Application of Microfabrication to Fluid Mechanics 197, 57-66, 1994.

8. Arkilic E. B., Breuer K. S. and Schmidt M. A., ‘Mass flow and tangential
momentum accommodation in silicon micro-machined channels’, J. Fluid
Mechanics, vol.437, pp.29-43, 2001.

9. Balakrishnan R., ‘An approach to entropy consistency in second-order


hydrodynamic equations’, J. Fluid Mech., vol. 503, pp. 201-245, 2004.

10. Barcelo B. T., ‘Electron beam measurements of the shock wave structure. Part 1,
Determination of the interaction potential of the noble gases from shock wave
structure experiments’, PhD thesis, California Institute of Technology, USA,
1971.
11. Bardow A. and Oettinger H. C., ‘Consequences of the Brenner modification to
the Navier- Stokes equations for dynamic light scattering’, Physica A, 373, p.88,
2007.

12. Bird G. A., ‘Aspects of the structure of strong shock waves’, Physics of Fluids,
13, p.1172, 1970.

13. Baysal O. and Aslan A. R., ‘Computing separated flows in MEMS devices’, In
Proceedings of ASME Fluids Engineering Division Summer Meeting, Montreal,
Quebec, Canada, Paper No. FEDSM2002-31157, July 14–18, 2002.

14. Bedeaux D., Kjelstrup S., Oettinger H. C., ‘On a possible difference between the
barycentric velocity and the velocity that gives translational momentum in fluids’,
Physica A., 371, pp.177-187, 2006.

15. Bejan A., Convection Heat Transfer, second ed., Wiley, New York, , Chapter 8,
p.293, 1995.

16. Beskok A., ‘Validation of a new velocity-slip model for separated gas micro-
flows’, Numerical Heat Transfer, Part B, 40:pp.451-471, 2001(a).

17. Beskok A., ‘Molecular based micro-fluidic simulation models’, In MEMS


Handbook, Gad-el-Hak M (ed.). CRC Press, New York, 8.1–8.28, 2001(b).

18. Bielenberg J.R. and Brenner H., ‘A continuum model of thermal transpiration’, J.
Fluid Mech., vol. 546, pp.1-23, 2006.

19. Bird R. B., Stewart W. E. and Lightfoot E. N., Transport Phenomena, Wiley, New
York, 1960.

20. Boltzmann L.: see C. Cercignani, Theory and Application of the Boltzmann
Equations, Elsevier, New York, 1975.

21. Brenner H., ‘Navier-Stokes revisited’, Physica A, 349, pp.60-132, 2005 (a).

22. Brenner H., ‘Kinematics of Volume Transport’, Physica A, 349, pp.11-59, 2005
(b).

196
23. Brenner H., ‘Nonisothermal Brownian Motion: Thermophoresis as the
Macroscopic Manifestation of Thermally Biased Molecular Motion’, Phys. Rev. E,
72, 061201, 2005 (c).

24. Brenner H. and Bielenberg J. R., ‘A continuum approach to phoretic motions:


Thermophoresis’, Physica A, vol. 335, Issue 2-4, pp.251-273, 2005 (d).

25. Brock J. R., ‘On the theory of thermal forces acting on aerosol particles, Journal
of colloid science, 17, pp. 768-780, 1962.

26. Celik B and Edis F. O., ‘Analysis of fluid flow through micro-fluidic devices using
characteristic-based-split procedure’, Int. J. Numer. Meth. Fluids, 51: pp.1041-
1057, 2006.

27. Celik B. And Edis F. O., ‘Computational Investigation of Micro Backward Facing
Step Duct Flow in Slip Regime’, Nanoscale and Microscale Thermophysical
Engineering, 11, pp.319-331, 2007.

28. Cercignani C., Frezzotti A. and Grosfils P., ‘The structure of an infinitely strong
shock wave’, Physics of Fluids, 11, p.2757, 1999.

29. Chakraborty, S., and Durst, F., ‘Derivations of Extended Navier-Stokes Equations
from Molecular Transport Considerations for Compressible Ideal Gas Flows:
Towards Extended Constitutive Forms’, Physics of Fluids, vol. 19, pp. 088104 (1-
4), 2007.

30. Chakraborty, S., ‘DSMC data of gas flows through microchannels with a backward
facing step’, unpublished data obtained through a private communication, 2010.

31. Chapman S. and Cowling T. G., The Mathematical Theory of Non-uniform Gases,
Cambridge University Press, Cambridge, 1953.

32. Chih-Ming Ho and Yu-Chong Tai, ‚Micro-Electro-Mechanical-Systems (MEMS)


and Fluid Flows‘, Annu. Rev. Fluid Mech., 30, 579-612, 1998.

33. Coelho R. M. L. and Silva Telles A., ‚Extended Graetz problem accompanied by
Dufour and Soret effects‘, Int. Journal of Heat and Mass Transfer, 45, pp. 3101-
3110, 2002.

34. Colin, S., Lalonde, P. and Caen, R., ‘Validation of a Second-Order Slip Flow
Model in Rectangular Microchannels’, Heat Transfer Eng., 25: 3 23-30, 2004.

35. Davis J. E., ‘Thermophoresis of Particles’, Encyclopedia of Surface and Colloid


Science, Edited by P. Somasundaran, 2nd Edition,Taylor and Francis, 2006.
197
36. Dongari, N., Agrawal, A. & Amit, A., ‘Analytical solution of gaseous slip flow in
long microchannels’, International Journal of Heat and Mass Transfer, 50, 3411-
3421, 2007.

37. Derjaguin B. V., Rabinovich YA. I., Storozhilova A. I. and Shcherbina G. I.,
‘Measurement of the Coefficient of Thermal Slip of Gases and the
Thermophoresis Velocity of Large-Size Aerosol Particles, Journal of Colloid and
Interface Science, Vol. 57, No. 3, pp. 451-461, 1976.

38. Dongari, N., Sharma, A. & Durst, F., ‚Pressure-driven diffusive gas flows in
microchannels: From the Knudsen to the continuum regimes‘, Microfluidics and
Nanofluidics, 6, pp. 679-692, 2009.

39. Dongari N., Sambasivam R. and Durst. F, ‘Extended Navier-Stokes equations and
treatments of microchannel gas flows’, JSME Journal of Fluid Science and
Technology, vol.4, No.2, pp.1-14, 2009.

40. Duan Z and Muzychka Y.S., ‘Slip flow in elliptic microchannels’, International
Journal of Thermal Sciences, 46, pp.1104-1111, 2007(a).

41. Duan Z and Muzychka Y.S., ‘Compressibility Effect on Slip Flow in Non-circular
Microchannels’, Nanoscale and Microscale Thermophysical Engineering, 11,
pp.259-272, 2007(b).

42. Durst F., Gomes J. and Sambasivam R., “Thermo-fluid-dynamics: Do we solve the
right kind of equations?”, 5th Symposium on Turbulence, Heat and Mass Transfer,
K. Hanjalic, Y. Nagamo & S. Jakirlic(ed.), vol.1, Dubrovnik, p.3, September 2006.

43. Durst F., “Theorie and Praxis der Stroemungsmechanik – eine Einfuehrung”,
2006 and ‘Fluid Mechanics – An Introduction to the Theory of Fluid Flows’,
Springer - Verlag, Heidelberg, 2008.

44. Edwards D. A., Brenner H. and Wasan D.T., ‚Interfacial Transport Processes and
Rheology‘, Butterworth-Heinemann, Boston, 1991.

45. Epstein P.S., ‚Zur Theorie des Radiometers‘, Z. Phys., 54, pp.537-563, 1929.

46. Errol, B., Arkilic, E. B., Kenneth, S., Breuer, K. S. & Schmidt M. A., ‘Mass flow
and tangential momentum accommodation in silicon micro-machined channels’,
J. Fluid Mech., 437, pp. 29-43, 2001.

198
47. Ezata K., Shehata A. M., Kunugi T. and McEligot D. M., ‘Numerical predictions
of transitional features of turbulent gas flows in circular tubes with strong
heating’, J. Heat transfer, 121, pp. 546-555, 1999.

48. Filimonov, D., Adachi, T., Sambasivam, R. and Durst, F., ‘Analytical treatments
of microchannel and micro-capillary flows, submitted for publication, 2011.

49. Fluent 6.3 Users guide manual, Fluent Inc., Lebanon, USA, 2008.

50. Gad-el-Hak M, ‘The Fluid Mechanics of Microdevices – The Freeman Scholar


Lecture’, Journal of Fluids Engineering, Vol.121, pp.5-33, March, 1999.

51. Gad-el-Hak M (ed.), MEMS Handbook, CRC Press, New York, 2002.

52. Garg V. K. and Jayaraj S., ‘Thermophoresis of aerosol particles in laminar flow
over inclined plates’, Int. J. Heat and Mass Transfer, Vol. 31, No. 4, pp. 875-890,
1988.

53. Greenshields C. J. and Reese J. M., ‘The Structure of Shock Waves as a Test of
Brenner’s Modifications to the Navier-Stokes Equations’, J. Fluid Mechanics,
vol.580, pp.407-429, 2007(a).

54. Greenshields, C.J. and Reese, J.M., ‘The structure of hypersonic shock waves
using Navier-Stokes equations modified to include mass diffusion’, In 2nd
European Conference on Aero-Space Sciences (EUCASS), Von Karman Institute,
2007(b).

55. Harley, J.C., Huang, Y., Bau, H. and Zemel, J.N., ‘Gas flow in microchannels’, J.
Fluid Mechanics, 284 257-274, 1995

56. Hirschfelder J. O., Curtiss C. F. and Bird R. B., ‘Molecular Theory of gases and
Liquids’, Wiley, New York, 1954.

57. Homsy G. M., Geyling F. T. and Walker K. L., ‘Blasius Series for Thermophoretic
Deposition of Small Particles’, Journal of Colloid and Interface Science, Vol. 83,
No. 2, pp. 495-501, 1981.

58. Karniadakis, G., Beskok, A. and Aluru, N., Microflows and Nanoflows -
Fundamentals and Simulation, Springer, 2005.

59. Kennard E. H., ‘Kinetic theory of gases’, McGraw Hill, New York, 1938.

199
60. Knudsen, M., ‘Die Gesetze der Molekularstroemung und der inneren
Reibungsstroemung der Gase durch Roehren’, Ann. Phys. vol. 28 pp 75-130, 1909.

61. Kursun U. and Kapat J. S., ‘Modelling of Microscale Gas Flows in Transition
Regime Part I: Flow Over Backward Facing Steps’, Nanoscale and Microscale
Thermophysical Engineering, 11, pp. 15-30, 2007.

62. Landau L. D. and Lifschitz E. M., Lehrbuch der Theoretischen Physik, Band VI.
Hydrodynamik, Akademie Verlag, Berlin, 1966.

63. Liu, J., Tai, Y.C., Pong, K. and Ho, C.M., ‘MEMS for pressure distribution studies
of gaseous flows in microchannels’, An Investigation of Micro Structures, Sensors,
Actuators, Machines and Systems. Proc. Ann. Int. Workshop MEMS, 8th,
Amsterdam, pp. 20915, New York: IEEE, 1995.

64. Lumpkin F. E. and Chapman D. R., ‘Accuracy of the Burnett equations for
hypersonic real gas flows’, AIAA Paper, 91-0771, 1991.

65. Macrossan M. N. and Lilley C. R., ‘Viscosity of argon at temperatures > 2000 K
from measured shock thickness’, Physics of Fluids, 15, p.3452, 2003.

66. Magyari E., ‘Thermophoretic deposition of particles in fully developed mixed


convection flow in a parallel plate vertical channel: the full analytical solution’,
Heat and Mass Transfer, 45, pp. 1473-1482, 2009.

67. Maitland, G. C. and Smith, E. B., ‘Critical reassessment of viscosities of 11


common gases’, J. Chem. Engng. Data, 17, p.150, 1972.

68. Maurer J., Tabeling P., Joseph P. and Willaime H, ‘Second-order slip laws in
microchannels for helium and nitrogen’, Physics of Fluids, Vol.15, No. 9,
pp.2613-2621, September 2003.

69. Maxwell, J., ‘On stress in rarefied gases arising from inequalities of temperature’,
Philos.Trans. R. Soc. London, 170, 231-256, 1879.

70. Mills A. F., Heat Transfer, R D Irwin Inc., chapter 5, p.395, 1992.

71. Mo G. and Rosenberger F., ‘Molecular dynamics simulation of flow in a two


dimensional channel with atomically rough walls’, Phys. Rev. A, 42, pp.4688-4692,
1990.

72. Navaneetha Krishnan, R., Filimonov, D., Sambasivam, R., Das, S. K. And Durst,
F., ‘Analytical and semi-analytical treatments of microchannel flows’, submitted
for publication, 2010.

200
73. Oettinger H. C., ‘Beyond Equilibrium Thermodynamics’, Wiley and Sons, 2005.

74. Pfahler, J., Harley, J.C., Bau, H. and Zemel J.N., ‘Gas and liquid flow in small
channels’, ASME DSC 32, pp.49-60, 1991.

75. Pong, K.C., Ho, C.M., Liu, J. and Tai, Y.C., ‘Nonlinear pressure distribution in
uniform microchannels’, ASME FED 197, pp.51-56, 1994.

76. Prandtl L., ‘The mechanics of viscous fluids’, In W. F. Durand (ed.), Aerodynamic
Theory, [1], pp.16-208, 1935.

77. Reed L. D. and Morrison F. A. Jr., ‘Motion of an Aerosol Particle in an Unsteady


Temperature Gradient’, Journal of Colloid and Interface Science, Vol. 42, No. 2,
pp. 358 – 365, 1973.

78. Reed L. D. and Morrison F. A. Jr., ‘Particle interactions in low Knudsen number
thermophoresis’, Journal of Aerosol Science, Vol.6, pp.349-365, 1975.

79. Reese J. M., ‘On the structure of shock waves in monatomic rarefied gases’, PhD
thesis, Oxford University, UK, 1993.

80. Reese J. M., Woods L. C., Thivet F. J. P. and Candel S. M., ‘A second-order
description of shock structure’, J. Comput. Phys., 117, pp.240-250, 1995.

81. Reif F., ‚Physikalische Statistik und Physik der Waerme‘, Walter De Gruyter and
Co., Berlin, 1987.

82. Reynolds O., ‘On certain dimensional properties of matter in gaseous state’, Phil.
Trans. Roy. Soc. (London), London, B170, pp.727-845, 1879.

83. Sambasivam R. and Durst F., Extended Fluid Mechanics Equations and their
Applications to Unsolved Flow Problems, Submitted for publication, 2011(a).

84. Sambasivam R. and Durst F., ‘Ideal gas flows through microchannels – Revisited’,
Microfluidics and Nanofluidics Handbook: Chemistry, Physics and Life Sciences
Principles, Edited by S. K. Mitra and S. Chakraborty, 2011(b).

85. Shehata A. M. and McEligot D. M., ‘Mean turbulence structure in the viscous layer
of strongly-heated internal gas flows: Measurements’, Int. J. Heat and Mass
Transfer, 41, pp.4297-4313, 1998.

86. Shih, J.C., Ho, C.M., Liu, J. and Tai, Y.C., ‘Nonlinear pressure distribution in
uniform microchannels’, ASME AMD-MD 238, 1995.

201
87. Shih, J.C., Ho, C.M., Liu, J. and Tai, Y.C., ‘Monoatomic and polyatomic gas
flow through uniform microchannels’, ASME DSC 59 197-203, 1996.

88. Steinhilper E. A., ‘Electron beam measurements of the shock wave structure. Part
1, The inference of intermolecular potentials from shock structure experiments’,
PhD thesis, California Institute of Technology, USA, 1972.

89. Sturtevant B. and Steinhilper E. A., ‘Intermolecular potentials from shock


structure experiments’, In 8th Intl. Symp. on Rarefied Gas Dynamics (ed. K.
Karamcheti), p.159, Academic, 1974.

90. Talbot L., ‘Thermophoresis - a review’, in: S.S. Fisher (Ed.), Rarefied Gas
Dynamics, Part 1, AIAA, New York, p. 467, 1981.

91. Tyndall J., ‘On dust and disease’, Proc. R. Inst. 6, pp.1-14, 1870.

92. White F. M., “Viscous Fluid Flow”, McGraw-Hill Book Company, New York,
1974.

93. Wu J. S and Lee F., ‘Pressure Boundary Treatment in Micromechanical Devices


Using The Direct Simulation Monte Carlo Method’, JSME International Journal,
Series B, vol. 44, No.3, pp.439-450, 2001.

94. Yang, Z. and Garimella, S. V., ‘Rarefied gas flow in microtubes at different inlet-
outlet pressure ratios’, Physics of Fluids, 21, 052005, 2009.

95. Zhang Y., Qin R. and Emerson D. R., ‘Lattice Boltzmann simulation of rarefied
gas flows in microchannels’, Physical Review E, 71, 047702, 2005.

96. Ziarani A. S. and Mohamad A. A, ‘Effect of wall roughness on the slip of fluid in a
microchannel’, Nanoscale and Microscale Thermophysical Engineering, 12,
pp.154-169, 2008.

97. Zienkiewicz O. C. and Codina R., ‘A general algorithm for compressible and
incompressible flow, part I, The split characteristic based scheme’, International
Journal for Numerical Methods in Fluids, 20, pp.869–885, 1995.

202
APPENDIX
In order to numerically simulate the extended Navier-Stokes equations proposed in this
thesis, the commercially available computational available Computational Fluid
Dynamics (CFD) code, FLUENT 6.3 and 12.1 was employed. FLUENT is an
unstructured, general purpose CFD-code, based on finite volume method. The classical
continuity, momentum and energy conservation equations, mentioned in equations
(2.1)- (2.3) and (2.25) are incorporated in this code. Therefore the additional terms,
arising out of the extended equations, were included in the code by writing suitable user
defined functions (UDF), as given below. Since the classical constitutive terms for the
diffusive transport of momentum and heat, mentioned in equations (2.24) and (2.26),
have already been included in the code, only the additional terms arising out of the
extensions were incorporated into the UDF functions.

UDFs employed in two dimensional wall bounded flows

The following User Defined Functions (UDF) were employed to simulate wall bounded
flows in the thesis such as gas flows through microchannels and capillaries, gas flow
through strongly heated pipes etc. The total velocity form of the extended equations,
given by equations (3.46), (3.49), (3.76) and (3.67), were used in the numerical
simulations of wall bounded flows. The functions named [Horizontal wall velocity] and
[Vertical wall velocity] were employed to specify the diffusion velocity at the horizontal
(along x1 direction) and vertical (along x 2 direction) walls, respectively. The other
additional terms in the left-hand sides of the momentum and energy equations, given by
equations (3.49) and (3.76), respectively, were estimated in the function named
[adjust_UDIs] and the volumetric source terms were added in the corresponding
momentum and energy equations employing the functions [x_mom_source],
[y_mom_source] and [energy_source], respectively. Further, the functions employed to
estimate the fluid properties such as thermal conductivity and to specify the constant
pressure and temperature boundary conditions at the inlet and outlet boundaries are also
provided here.

#include "udf.h"

DEFINE_PROPERTY (thermal_conductivity, cell, thread)


{
real K, mu, Cp;
cell_t t;
mu = C_MU_L(cell,thread);
Cp = C_CP(cell,thread);
K = mu*Cp/0.72;
return K;
}
DEFINE_PROFILE (inlet_pressure, t, position)
{
real rho, vel;
face_t f;
cell_t c0;
Thread *tc0;

begin_f_loop(f,t)
{
c0 = F_C0(f,t);
tc0 = THREAD_T0(t);
rho = C_R(c0,tc0);
vel = F_U(f,t);
{
if (F_U(f,t) >= 0.0)
{
F_PROFILE(f,t,position) = 624014.0+(0.5*rho*vel*vel);
}
else
F_PROFILE(f,t,position) = 624014.0;
}
}
end_f_loop(f,t)
}

DEFINE_PROFILE (inlet_temperature, t, position)


{
real cp, vel,Ma,T;
face_t f;
cp=1040.67;

begin_f_loop(f,t)
{
vel=F_U(f,t);
T=F_T(f,t);
Ma=vel/sqrt(1.4*8314.5*T/28.0134);
{
if (vel >= 0.0)
F_PROFILE(f,t,position) = 330.0+(0.5*F_U(f,t)*F_U(f,t)/1040.67);
else
F_PROFILE(f,t,position) = 330.0;
}
}
end_f_loop(f,t)
}
204
DEFINE_PROFILE (outlet_pressure, t, position)
{
real rho, vel,u,v;
face_t f;

begin_f_loop(f,t)
{
rho = F_R(f,t);
u = F_U(f,t);
v = F_V(f,t);
vel = sqrt((u*u)+(v*v));
{
if (u <= 0.0)
{
F_PROFILE(f,t,position) = 271983+ (0.5*rho*vel*vel) ;
}
else
F_PROFILE(f,t,position) =271983;
}
}
end_f_loop(f,t)
}

DEFINE_PROFILE (horizontal_wall_velocity, tf, pos)


{
face_t f;
cell_t c0;
Thread *tc0;
real press,in_press,vis,p_grad,rho,T,T_grad;

begin_f_loop(f,tf)
{
c0 = F_C0(f,tf);
tc0 = THREAD_T0(tf);
press = C_P(c0,tc0);
p_grad=C_P_G(c0,tc0)[0];
in_press = 1.0/press;
vis = C_MU_L(c0,tc0);
T=C_T(c0,tc0);
rho=C_R(c0,tc0);
T_grad=C_T_G(c0,tc0)[0];
F_PROFILE(f,tf,pos)=-1.0*vis*((in_press*p_grad)-((0.5/T)*T_grad))/rho;
}
end_f_loop(f,tf)
}
205
DEFINE_PROFILE (vertical_wall_velocity, tf, pos)
{
face_t f;
cell_t c0;
Thread *tc0;
real press,in_press,vis,p_grad,rho,T,T_grad;

begin_f_loop(f,tf)
{
c0 = F_C0(f,tf);
tc0 = THREAD_T0(tf);
press = C_P(c0,tc0);
p_grad = C_P_G(c0,tc0)[1];
in_press = 1.0/press;
vis = C_MU_L(c0,tc0);
T = C_T(c0,tc0);
rho = C_R(c0,tc0);
T_grad=C_T_G(c0,tc0)[1];
F_PROFILE(f,tf,pos)=-1.0*vis*((in_press*p_grad)-((0.5/T)*T_grad))/rho;
}
end_f_loop(f,tf)
}

DEFINE_ADJUST (adjust_UDSIs,d)
{
real vis,u,v,temp1,in_temp1,in_press,press,rho,in_rho,x_pos,sp_heat,tau_c_11,
tau_c_12, tau_c_22, x[ND_ND], T_ref=298.15;
Thread *t;
cell_t c;

thread_loop_c(t,d)
{
begin_c_loop(c,t)
{
C_CENTROID(x,c,t);
x_pos = x[0];
vis = C_MU_L(c,t);
sp_heat = C_CP(c,t);
rho = C_R(c,t);
temp1 = C_T(c,t);
press = C_P(c,t);
u = C_U(c,t);
v = C_V(c,t);
x_pos = x[0];
in_press = 1.0/press;
206
C_UDSI(c,t,0) = - vis*in_press*C_P_G(c,t)[0];
C_UDSI(c,t,1) = - vis*in_press*C_P_G(c,t)[1];
C_UDSI(c,t,2) = C_UDSI(c,t,0)/rho;
C_UDSI(c,t,3) = C_UDSI(c,t,1)/rho;
C_UDSI(c,t,4)=(2.0/3.0)*vis*((2.0*C_UDSI_G(c,t,2)[0])-
C_UDSI_G(c,t,3)[1])+C_UDSI(c,t,0)*C_UDSI(c,t,2))-
((2.0/3.0)*C_UDSI(c,t,0)*C_U(c,t));
C_UDSI(c,t,5)=vis*(C_UDSI_G(c,t,3)[0]+C_UDSI_G(c,t,2)[1])-
(C_UDSI(c,t,1)*C_UDSI(c,t,2));
C_UDSI(c,t,6)=(2.0/3.0)*vis*((2.0*C_UDSI_G(c,t,3)[1])-
C_UDSI_G(c,t,2)[0])-(C_UDSI(c,t,1)*C_UDSI(c,t,3))-
((2.0/3.0)*C_UDSI(c,t,1)*C_V(c,t));
C_UDSI(c,t,7)=C_U(c,t)-C_UDSI(c,t,2);
C_UDSI(c,t,8)=C_V(c,t)-C_UDSI(c,t,3);
tau_c_11=((4.0/3.0)*vis*C_UDSI_G(c,t,7)[0])-((2.0/3.0)*vis*
C_UDSI_G(c,t,8)[1]);
tau_c_12=vis*(C_UDSI_G(c,t,7)[1]+C_UDSI_G(c,t,8)[0]);
tau_c_22=((4.0/3.0)*vis*C_UDSI_G(c,t,8)[1])-((2.0/3.0)*vis*
C_UDSI_G(c,t,7)[0]);
C_UDSI(c,t,9)=tau_c_11+(C_UDSI(c,t,0)*C_UDSI(c,t,2))+((2.0/3.0)*
((C_UDSI(c,t,0)*C_UDSI(c,t,7))+
(C_UDSI(c,t,1)*C_UDSI(c,t,8))));
C_UDSI(c,t,10)=tau_c_12+(C_UDSI(c,t,1)*C_UDSI(c,t,7));
C_UDSI(c,t,11)=tau_c_12+(C_UDSI(c,t,0)*C_UDSI(c,t,8));
C_UDSI(c,t,12)=tau_c_22+(C_UDSI(c,t,1)*C_UDSI(c,t,3))+((2.0/3.0)*
((C_UDSI(c,t,0)*C_UDSI(c,t,7))+(C_UDSI(c,t,1)*C_UDSI(c,t,8))));
}
end_c_loop(c,t)
}
}

DEFINE_SOURCE (x_mom_source, c, t, dS, eqn)


{
real source,x_pos, x[ND_ND];
C_CENTROID(x,c,t);
x_pos=x[0];
source=-C_UDSI_G(c,t,4)[0]-C_UDSI_G(c,t,5)[1];
dS[eqn]=0.0;
return source;
}

DEFINE_SOURCE (y_mom_source, c, t, dS, eqn)


{
real source,x_pos, x[ND_ND];
C_CENTROID(x,c,t);
207
x_pos=x[0];
source=-C_UDSI_G(c,t,5)[0]-C_UDSI_G(c,t,6)[1];
dS[eqn]=0.0;
return source;
}

DEFINE_SOURCE (energy_source, c, t, dS, eqn)


{
real source,x_pos,Vsource, x[ND_ND];
C_CENTROID(x,c,t);
x_pos=x[0];
Vsource=((C_MU_L(c,t)*((4.0*C_U_G(c,t)[0]/3.0)-(2.0*C_V_G(c,t)[1]/3.0)))*
C_UDSI_G(c,t,7)[0])+((C_MU_L(c,t)*(C_U_G(c,t)[1]+C_V_G(c,t)[0]))
*(C_UDSI_G(c,t,7)[1]+C_UDSI_G(c,t,8)[0]))+((C_MU_L(c,t)*
((4.0*C_V_G(c,t)[1]/3.0)-(2.0*C_U_G(c,t)[0]/3.0)))*C_UDSI_G(c,t,8)[1]);
source=(C_P(c,t)*(C_UDSI_G(c,t,2)[0]+C_UDSI_G(c,t,3)[1]))+(C_U(c,t)*
(C_UDSI_G(c,t,9)[0]+C_UDSI_G(c,t,10)[1]))+
(C_V(c,t)*(C_UDSI_G(c,t,11)[0]+C_UDSI_G(c,t,12)[1]))+Vsource;
dS[eqn]=0.0;
return source;
}

UDFs employed in shock wave simulations

The following User Defined Functions (UDF) were employed to simulate shock waves
employing the extended equations, given by equations (3.21), (3.37), (3.38), (3.40) and
(3.64). The other extended terms in the left-hand sides of the continuity, momentum
and energy equations were estimated in the function named [adjust_UDIs] and the
volumetric source terms were added in the continuity, momentum and energy equations
employing the functions [mass_source], [mom_source] and [energy_source],
respectively. Further, the functions employed to estimate the fluid properties such as
thermal conductivity and to specify the constant pressure and temperature boundary
conditions at the inlet and outlet boundaries are also provided here.

#include "udf.h"

DEFINE_PROPERTY (thermal_conductivity, cell, thread)


{
real K, mu, Cp;
cell_t c;
mu = C_MU_L(cell,thread);
Cp = C_CP(cell,thread);
K=mu*Cp/0.667;
return K;
}
208
DEFINE_PROFILE (outlet_pressure, t, position)
{
real rho, vel;
face_t f;
begin_f_loop(f,t)
{
rho = F_R(f,t);
vel = F_U(f,t);
if (vel <= 0.0)
{
F_PROFILE(f,t,position) = 1006.42 + (0.5*rho*vel*vel) ;
}
else
F_PROFILE(f,t,position) = 1006.42;
}
end_f_loop(f,t)
}

DEFINE_PROFILE (outlet_temperature, t, position)


{
real vel;
face_t f;
begin_f_loop(f,t)
{
vel = F_U(f,t);
if (vel <= 0.0)
{
F_PROFILE(f,t,position) = 11605.8 + (0.5*vel*vel/520.64) ;
}
else
F_PROFILE(f,t,position) = 11605.8;
}
end_f_loop(f,t)
}

DEFINE_ADJUST (adjust_UDSIs,d)
{
real vis,u,temp1,in_temp1,in_rho,rho,x_pos,dRdx,dTdx,sp_heat, x[ND_ND];
real T_ref=298.15;
Thread *t;
cell_t c;
thread_loop_c(t,d)
{
begin_c_loop(c,t)

209
{
C_CENTROID(x,c,t);
x_pos=x[0];
vis=C_MU_L(c,t);
sp_heat=C_CP(c,t);
rho=C_R(c,t);
temp1=C_T(c,t);
u=C_U(c,t);
dRdx=C_R_RG(c,t)[0];
dTdx=C_T_G(c,t)[0];
if (dTdx<0.0) dTdx=0.0;
if (dRdx<0.0) dRdx=0.0;
in_temp1 = 1.0/temp1;
in_rho = 1.0/rho;
C_UDSI(c,t,0)=-vis*((in_rho*dRdx)+(0.5*in_temp1*dTdx));
C_UDSI(c,t,1)=C_UDSI(c,t,0)/rho;
C_UDSI(c,t,2)=(-4.0*C_UDSI(c,t,0)*u/3.0);
C_UDSI(c,t,3)=-(C_UDSI(c,t,0)*(sp_heat/1.667)*(temp1-T_ref))+
(C_UDSI(c,t,2)*u);
}
end_c_loop(c,t)
}
}

DEFINE_SOURCE (mass_source, c, t, dS, eqn)


{
real source,rho, x_pos, x[ND_ND];
C_CENTROID(x,c,t);
x_pos=x[0];
rho=C_R(c,t);
source=-C_UDSI_G(c,t,0)[0];
dS[eqn]=0.0;
return source;
}

DEFINE_SOURCE (mom_source, c, t, dS, eqn)


{
real source,u,x_pos, x[ND_ND];
C_CENTROID(x,c,t);
x_pos=x[0];
u=C_U(c,t);
source=C_UDSI_G(c,t,2)[0];
dS[eqn]=0.0;
return source;
}
210
DEFINE_SOURCE (energy_source, c, t, dS, eqn)
{
real source,x_pos, x[ND_ND];
C_CENTROID(x,c,t);
x_pos=x[0];
source=C_UDSI_G(c,t,3)[0];
dS[eqn]=0.0;
return source;
}

DEFINE_PROFILE (inlet_pressure, t, position)


{
real rho, vel;
face_t f;
begin_f_loop(f,t)
{
rho = F_R(f,t);
vel = F_U(f,t);
if (vel >= 0.0)
{
F_PROFILE(f,t,position) = 6.665 + (0.5*rho*vel*vel) ;
}
else
F_PROFILE(f,t,position) = 6.665;
}
end_f_loop(f,t)
}

DEFINE_PROFILE (inlet_temperature, t, position)


{
real vel;
face_t f;
begin_f_loop(f,t)
{
vel = F_U(f,t);
if (vel >= 0.0)
{
F_PROFILE(f,t,position) = 300.0 + (0.5*vel*vel/520.64) ;
}
else
F_PROFILE(f,t,position) = 300.0;
}
end_f_loop(f,t)
}

211
212
Erweiterte Navier-Stokes-Gleichungen:
Ableitungen und Anwendungen auf
StrÖmungsprobleme

Der Technischen Fakultät der


Universität Erlangen-Nürnberg
zur Erlangung des Grades

DOKTOR-INGENIEUR

vorgelegt von

Rajamani Sambasivam

Erlangen, 2013
214
Zusammenfassung
Die vorliegende Arbeit fasst die von Prof. Franz Durst, emeritierter Professor der
Universität Erlangen-Nürnberg, betreute Forschungsarbeit des Autors im Bereich der
erweiterten Navier-Stokes-Gleichungen zusammen. Im Rahmen einer früheren
Zusammenarbeit erkannte der Autor, dass die Beschreibung durch die klassischen
Navier-Stokes-Gleichungen unter bestimmten Strömungsbedingungen nicht durch die
entsprechenden experimentellen Messergebnisse bestätigt wurde. Dies erstaunte den
Autor und führte zu einer nähergehenden Betrachtung der Ableitungen der
Transportgleichungen von dem ersten Prinzip. Der Autor stellte fest, dass sich die
Ableitungen der klassischen Gleichungen als fehlerhaft erwiesen, wenn sie angewandt
wurden, um Gasströmungen mit hohen Dichte-, Druck- und Temperaturgradienten, die
zu zusätzlichen diffusiven Massenströmungen führten, zu lösen. Es wurde argumentiert,
dass dieser zusätzliche Massenstrom den konvektiven Massenströmen, die in den
klassischen Navier-Stokes Gleichungen behandelt wurden, hinzugefügt werden
mussten. Außerdem war es auch erforderlich, die abgewandelten konstitutiven
Beziehungen des Molekulartransports von Impuls und Wärme in die Gleichungen zu
integrieren. Daher kann argumentiert werden, dass die klassischen Navier-Stokes-
Gleichungen nur Gültigkeit besitzen, wenn es in idealen Gasströmungen keine Dichte-
und Temperaturgradienten gibt.

Die vorliegende Arbeit beginnt mit einer Zusammenfassung der Entwicklung der
klassischen Navier-Stokes-Gleichungen, die auf dem bestehenden, historischen Wissen
der Strömungsmechanik basiert. Ein kurzer geschichtlicher Abriss der Entwicklung der
klassischen Navier-Stokes-Gleichungen, die die Kontinuitätsgleichung, die
Impulsgleichung und die Energiegleichung umfassen, wird aufgezeigt. Dabei wird
jedoch auf die jüngsten Forschungsentwicklungen eingegangen und der Wissensstand
zu Beginn der eigenen Ableitungen der grundlegenden Strömungsgleichungen bei in
idealen Gasströmungen vorkommenden Dichte- und Temperaturgradienten
zusammengefasst. Diese Ableitungen führten zu einem neuen Gleichungssystem, auch
als erweiterte Navier-Stokes Gleichungen bezeichnet. Die in dieser Arbeit abgeleiteten
Gleichungen wurden mit anderen erweiterten Strömungsgleichungen verglichen, die auf
völlig unterschiedlichen Betrachtungen beruhten, um Unterschiede zwischen den
klassischen und den erweiterten Gleichungen festzustellen.

Anschließend wurden unter Anwendung der erweiterten Navier-Stokes-Gleichungen,


die in dieser Arbeit abgeleitet wurden, Untersuchungen von idealen Gasströmungen
durch Mikrokanäle und Kapillare durchgeführt. Die erste Arbeit des Autors befasste
sich mit numerischen Vorhersagen der Gasströmungen durch Mikrokanäle, und es
wurde, ohne Bezug auf die Maxwell-Slip Randbedingung an den festen Wänden, eine
hervorragende Übereinstimmung mit den experimentellen Messergebnissen erzielt.
Basierend auf den Erkenntnissen aus den numerischen Simulationen half er schließlich
auch einer Gruppe von Wissenschaftlern, eine vollständige, analytische Lösung der
215
Gasströmungen durch Mikrokanäle und Kapillare abzuleiten. Aufgrund dieser
analytischen Lösungsmethode ist es nicht nur möglich, eine genaue Beschreibung der
Geschwindigkeitsverteilung, sondern auch der Druckverteilung innerhalb des
Mikrokanals zu erhalten. Noch wichtiger ist, dass ein charakteristischer Druck,
basierend auf Geometrie und physikalische Eigenschaften des Gases, bestimmt wurde,
der die Eigenschaften der Strömung genau beschreibt. Außerdem konnten mit Hilfe der
erweiterten Navier-Stokes-Gleichungen genaue numerische Beschreibungen der
Strömungen durch Mikrokanäle mit Trennung zur Verfügung gestellt werden. Eine
schrittweise rückwärts fließende Strömung wurde als Beispiel gewählt, um die
Zweckmäßigkeit der erweiterten Gleichungen für die Vorhersage der Gasströmungen
mit Trennung zu zeigen, und hervorragende Übereinstimmungen wurden mit DSMC-
Simulationen erzielt.

Nachfolgend wurden numerische Berechnungen für Gasströmungen ausgeführt, die sich


durch hohe Temperaturgradienten auszeichnen. Die Berechnung eindimensionaler
Überschall- und Hyperschallströmungen von Verdichtungsstößen wurden ausgewählt,
um die Nützlichkeit der erweiterten Gleichungen zu demonstrieren. Hervorragende
Übereinstimmung der berechneten Strukturen der Verdichtungsstöße mit
experimentellen Ergebnissen wurde erhalten. Vergleiche verschiedenster Eigenschaften,
wie die inverse Dicke der Dichteverteilung und deren Asymmetrie sowie die
Temperatur-Dichte-Aufteilung, wurden durch die Berechnungen mit den erweiterten
Navier-Stokes-Gleichungen gleichfalls bereitstellt. Eine gute Übereinstimmung mit
Experimenten wurde erhalten.

Es war interessant, und dies ist in der Dissertation erläutert, dass es gelang, unter
Anwendung der Navier-Stokes-Gleichungen eine Reihe von ungelösten Problemen, wie
Thermophorese und thermische Transpiration, physikalisch sinnvoll zu erklären. Weiter
wurde evident demonstriert, dass man die erweiterten Gleichungen auch bei bestimmten
Gasströmungsproblemen zur Anwendung bringen muss, die große geometrische
Abmessungen besitzen. Ohne diese Gleichungen erhält man keine genauen und
detaillierten Beschreibungen der Strömung und der Wärmeübertragung. Es wird ferner
die Notwendigkeit herausgestellt, dass weitere Forschungsarbeiten notwendig sind, um
die Möglichkeiten der erweiterten Navier-Stokes-Gleichungen herauszustellen.

216
CURRICULUM VITAE

Name Rajamani Sambasivam

Date of Birth 11th July 1975

Place of Birth Tirunelveli, Tamilnadu, India

Nationality Indian

Marital Status Married to Debjani

No. of children 2

1981 – 1990 Primary and High School in Tirunelveli, Tamilnadu, India

1991 – 1992 Higher Secondary School in Tirunelveli, Tamilnadu, India

1992 – 1996 Bachelor of Engineering (Mech. Engg.), Government


College of Engineering, Tirunelveli, Tamilnadu, India

1996 – 1999 Assistant Engineer – Plant Engg., M/s. Asian Paints (I) Ltd.,
Penta Division, Cuddalore, Tamilnadu, India

1999 – 2001 M. S. (Mechanical Engineering), Specialisation – Thermal


Engg., Indian Institute of Technology, Kharagpur, India

2001 – 2009 Researcher, Research and Development, TATA Steel,


Jamshedpur, India

2004 – 2005 Guest Researcher, Institute of Fluid Mechanics (LSTM),


University of Erlangen, Germany

2009 – Till date Head – Process Modelling and Visualisation, Technology


Group (Global Wires and Longs), TATA Steel, Jamshedpur,
India

217

You might also like