Unsteady Combustor Physics 2nd Edition Tim C. Lieuwen download
Unsteady Combustor Physics 2nd Edition Tim C. Lieuwen download
https://ebookmeta.com/product/unsteady-combustor-physics-2nd-
edition-tim-c-lieuwen/
https://ebookmeta.com/product/practical-physics-for-college-
students-2nd-edition-tim-l-ross/
https://ebookmeta.com/product/physics-of-nonlinear-optics-2nd-
edition-yvgs-vijayan-c-murti/
https://ebookmeta.com/product/handbook-of-radiotherapy-
physics-2nd-edition-philip-mayles-alan-e-nahum-j-c-rosenwald/
https://ebookmeta.com/product/sql-server-analytical-toolkit-
using-windowing-analytical-ranking-and-aggregate-functions-for-
data-and-statistical-analysis-1st-edition-angelo-bobak/
Futanari in Deep Space Julie Law
https://ebookmeta.com/product/futanari-in-deep-space-julie-law/
https://ebookmeta.com/product/frontiers-in-clinical-drug-
research-cns-and-neurological-disorders-volume-4-1st-edition-
atta-ur-rahman/
https://ebookmeta.com/product/sir-isaac-newton-s-mathematical-
principles-on-natural-philosophy-and-his-system-of-the-world-
volume-2-the-system-of-the-world/
https://ebookmeta.com/product/making-black-scientists-a-call-to-
action-1st-edition-marybeth-gasman/
https://ebookmeta.com/product/exploring-innovation-in-a-digital-
world-cultural-and-organizational-challenges-1st-edition-
federica-ceci/
The Ultimate Guide to SAT Grammar by Erica Meltzer
digital sat 6th Edition Erica L. Meltzer
https://ebookmeta.com/product/the-ultimate-guide-to-sat-grammar-
by-erica-meltzer-digital-sat-6th-edition-erica-l-meltzer/
Unsteady Combustor Physics
Second Edition
Tim C. Lieuwen is Regents’ Professor and Executive Director of the Strategic Energy
Institute at Georgia Tech. He is also the founder and CTO of TurbineLogic, an energy
analytics firm. He has authored 4 books and over 350 other publications. Board
positions include goveming/advisory boards for Oak Ridge National Lab, Pacific
Northwest National Lab, and the National Renewable Energy Lab, and appointment
by the DOE Secretary to the National Petroleum Counsel. He is an elected member of
the National Academy of Engineering, a fellow of ASME and AIAA, and recipient of
the AIAA Lawrence Sperry Award and ASME’s George Westinghouse Gold Medal.
Unsteady Combustor Physics
Second Edition
TIM C. LIEUWEN
Georgia Institute o f Technology
C a m b r id g e
U N IV E R SIT Y PR ESS
C a m b r id g e
U N IV E R S IT Y P R E SS
314-321, 3rd Floor, Plot 3, Splendor Forum, Jasola District Centre, New Delhi - 110025, India
103 Penang Road, #05-06/07, Visioncrest Commercial, Singapore 238467
www.cambridge.org
Information on this title: www.cambridge.org/9781108841313
DOI: 10.1017/9781108889001
© Tim C. Lieuwen 2012, 2021
This publication is in copyright. Subject to statutory exception
and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without the written
permission of Cambridge University Press.
First published 2012
Printed in the United Kingdom by TJ Books Limited, Padstow, Cornwall
A catalogue record for this publication is available from the British Library.
Introduction 1
1 Basic Equations 9
6 Acoustic Wave Propagation II: Heat Release, Complex Geometry, and Mean
Flow Effects 210
8 Ignition 296
Index 511
Detailed Contents
Introduction 1
Updates to the Second Edition 4
1 Basic Equations 9
1.1 Thermodynamic Relations in a Multicomponent Perfect Gas 9
1.2 Continuity Equation 10
1.3 Momentum Equation 11
1.4 Species Conservation Equation 12
1.5 Energy Equation 13
1.6 Aside: Discussion of the Vorticity and Circulation Equations 17
1.7 Nomenclature 20
1.7.1 Latin Alphabet 21
1.7.2 Greek Alphabet 23
1.7.3 Subscripts 24
1.7.4 Superscripts 25
1.7.5 Other Symbols 25
Exercises 25
References 26
8 Ignition 296
8.1 Overview 296
8.2 Autoignition 298
8.2.1 Ignition of Homogeneous, Premixed Reactants 298
8.2.2 Effects of Losses and Flow Inhomogeneity 301
8.3 Forced Ignition 311
Exercises 316
References 317
Index 511
Acknowledgments
Many individuals must be acknowledged for the completion of this book. First, I am
deeply appreciative to my dear wife, Rinda, and daughters Liske, Anneke, Carolina,
and Janna Lieu wen for their love, encouragement, and support.
This book would not have been possible without the financial support provided
through Joseph Citeno, which got the project kicked off, and the support of Vigor
Yang through my department. I am deeply grateful for their support, which made
initiating this project possible.
Next, this book would never have been possible without the enormous help
provided by my group here at Georgia Tech. They were a great help in pulling
together references, performing calculations, critiquing arguments, fleshing out deriv
ations, catching mistakes, and being a general sounding board. Particular thanks go to
Mike Aguilar, Alberto Amato, Ianko Chterov, Jack Crawford, Ben Emerson, Chris
Foley, Julia Lundrigan, Nick Magina, Mike Malanoski, Andrew Marshall, Jacqueline
O’Connor, Shreekrishna, Vishal Acharya Srinivas, Dong-Hyuk Shin, Ryan Sullivan,
Prabhakar Venkateswaran, and Ben Wilde. I have been very fortunate to have had
such a great team to work with and I thank all of them for their help.
Next, special thanks to Ben Bellows, Enrique Portillo Bilbao, Baki Cetegen, Jeff
Cohen, Joel Daou, Catalin Fotache, Fei Han, Santosh Hemchandra, Hong Im,
Matthew Juniper, Vince McDonell, Randal McKinney, Venkat Narra, Bobby Noble,
Preetham, Rajesh Rajaram, Mike Renfro, Paul Ronney, Dom Santavicca, Thomas
Sattelmayer, David Scarborough, Thierry Schuller, Santosh Shanbhogue, Shiva
Srinivasan, R.I. Sujith, Sai Kumar Thumuluru, and Qingguo Zhang for their feedback
and suggestions on the outline and content. In addition, Siva Harikumar, Faisal
Ahmed, and Jordan Blimbaum were a great editorial support team.
In addition, my sincere thanks go to my colleagues and mentors Ben Zinn, Robert
Loewy, Lakshmi Sankar, Jeff Jagoda, Jerry Seitzman, Suresh Menon, and Vigor Yang
for their help and support.
I am deeply appreciative of the suggestions for second edition contents and
comments on the manuscript by many of my colleagues, including Ben Emerson,
Vishal Acharya, Jacqueline O’Connor, Santosh Hemchandra, and Hong Im.
I am also very thankful for the help and support of Austin Matthews in helping to
pull together this second edition. His carefulness and attention to detail made this a
much better book.
Acknowledgments xv
This book is about unsteady combusting flows, with a particular emphasis on the
system dynamics that occur at the intersection of the combustion, fluid mechanics, and
acoustic disciplines - i.e„ on combustor physics. In other words, this is not a
combustion book - rather, it treats the interactions of flames with unsteady flow
processes that control the behavior of combustor systems. While numerous topics in
reactive flow dynamics are “unsteady” (e.g., internal combustion engines, detonations,
flame flickering in buoyancy-dominated flows, thermoacoustic instabilities), this text
specifically focuses on unsteady combustor issues in high Reynolds number, gas-
phase flows. This book is written for individuals with a background in fluid mechanics
and combustion (it does not presuppose a background in acoustics), and is organized
to synthesize these fields into a coherent understanding of the intrinsically unsteady
processes in combustors.
This book follows several texts or monographs which have treated related topics -
including Toong’s Combustion Dynamics [1], Crocco and Cheng’s Theory o f
Combustion Instability in Liquid Propellant Rocket Motors [2], Liquid Propellant
Rocket Combustion Instability by Harrje and Reardon [3], Putnam’s Combustion
Driven Oscillations in Industry [4], Fred Culick’s Unsteady Motions in Combustion
Chambers fo r Propulsion Systems [5], or Lieuwen and Yang’s edited Combustion
Instabilities in Gas Turbine Engines [6], Similarly, several dedicated texts on turbulent
combustion have been written, including Peters [7] and Lipatnikov [8],
Unsteady combustor processes define many of the most important considerations
associated with modem combustor design. These unsteady processes include transi
ent, time harmonic, and stochastic processes. For example, ignition, flame blowoff
and flashback are transient combustor issues that often define the range of fuel/air
ratios or velocities over which a combustor can operate. As we discuss in this book,
these transient processes involve the coupling of chemical kinetics, mass and energy
transport, flame propagation in high shear flow regions, hydrodynamic flow stability,
and interaction of flame-induced dilatation on the flow field - much more than a
simple balance of flame speed and flow velocity.
Similarly, combustion instabilities are a time-harmonic unsteady combustor issue
where the unsteady heat release excites natural acoustic modes of the combustion
chamber. These instabilities cause such severe vibrations in the system that they can
impose additional constraints on where combustor systems can be operated. The
acoustic oscillations associated with these instabilities are controlled by the entire
2 Introduction
combustor system; i.e., they involve the natural acoustics of the coupled plenum, fuel
delivery system, combustor, and turbine transition section. Moreover, these acoustic
oscillations often excite natural hydrodynamic instabilities of the flow, which then
wrinkle the flame front and cause modulation of the heat release rate. As such,
combustion instability problems involve the coupling of acoustics, flame dynamics,
and hydrodynamic flow stability.
Turbulent combustion itself is an intrinsically unsteady problem involving stochas
tic fluctuations that are both stationary (such as turbulent velocity fluctuations) and
nonstationary (such as turbulent flame brush development in attached flames).
Problems such as turbulent combustion noise generation require an understanding of
the broadband fluctuations in heat release induced by the turbulent flow, as well as the
conversion of these fluctuations into propagating sound waves. Moreover, the turbu
lent combustion problem is a good example for a wider motivation of this book -
many time-averaged characteristics of combustor systems cannot be understood
without understanding their unsteady features. For example, the turbulent flame speed,
related to the time-averaged consumption rate of fuel, can be one to two orders of
magnitude larger than the laminar flame speed, precisely because of the effect of
unsteadiness on the time-averaged burning rate. In turn, crucial issues such as flame
spreading angle and flame length, which then directly feed into basic design consider
ations such as locations of high combustor wall heat transfer, or combustor length
requirements, are then directly controlled by unsteadiness.
Even in nonreacting flows, intrinsically unsteady flow dynamics control many
time-averaged flow features. For example, it became clear a few decades ago that
turbulent mixing layers did not simply consist of broadband turbulent fluctuations, but
were, rather, dominated by quasi-periodic structures. Understanding the dynamics of
these large-scale structures has played a key role in our understanding of the time-
averaged features of shear layers, such as growth rates, mixing rates, or exothermicity
effects. Additionally, this understanding has been indispensable in understanding
intrinsically unsteady problems, such as how shear layers respond to external forcing.
Similarly, many of the flow fields in combustor geometries are controlled by
hydrodynamic flow instabilities and unsteady large-scale structures that, in turn, are
also profoundly influenced by combustion-induced heat release. It is well known that
the instantaneous and time-averaged flame shapes and recirculating flow fields in
many combustor geometries often bear little resemblance to each other, with the
instantaneous flow field exhibiting substantially more flow structures and asymmetry.
Flows with high levels of swirl are a good example of this, as shown by the
comparison of time-averaged (a) and instantaneous (b-d) streamlines in Figure 1.1.
Understanding such features as recirculation zone lengths and flow topology, and how
these features are influenced by exothermicity or operational conditions, necessarily
requires an understanding of the dynamic flow features. To summarize, continued
progress in predicting steady-state combustor processes will come from a fuller
understanding of their time dynamics.
Modern computations and diagnostics have revolutionized our understanding of the
spatiotemporal dynamics of flames since the publication of Markstein’s Nonsteady
Introduction 3
Figure 1.1 (a) Time-averaged and (b-d) instantaneous flow field in a swirling combustor flow.
Dashed line denotes isocontour of zero axial velocity and shaded regions denote vorticity
values. Image courtesy of M. Aguilar, M. Malanoski, and J. O ’Connor.
Figure 1.2 Line-of-sight (top) and planar (bottom) OH-PLIF images of turbulent, swirling
flame [10]. Images courtesy of B. Bellows.
It is hard to believe that 10 years have gone by since we started this project. Many of
the motivators and drivers of this book remain the same, but much has changed as
well. From a societal point of view, the march toward decarbonization is accelerating,
motivating topics like hydrogen combustion or combustor operability limits of alter
native fuels. The commercial space market has taken off and combustion instabilities
remain a key risk for rocket development. Major interest has developed in rotating
detonation engines, where the wave dynamics in annular passages and coupled
injector dynamics mirror many similar combustion instability topics. There is a
resurgence of interest in data-driven approaches for the analysis of complex data sets,
active control, or prediction of future or unmeasured system behaviors. Significant
developments have also taken place in hydrodynamic stability, particularly
reacting flows.
With this in mind, this book has been refreshed and updated. New sections have
been added or major updates made in the following sections:
References
[1] Toong T.Y., Combustion Dynamics: The Dynamics o f Chemically Reacting Fluids. 1983,
McGraw-Hill.
[2] Crocco L. and Cheng S.I., Tlieoiy o f Combustion Instability in Liquid Propellant Rocket
Motors. AGARDograph No. 8. 1956, Butterworths Scientific Publications.
[3] Harrje D.T. and Reardon F.H., Liquid Propellant Rocket Combustion Instability. 1972,
NASA.
[4] Putnam A.A., Combustion-Driven Oscillations in Industry. Fuel and Energy Science
Series. 1971, American Elsevier Publishing Company, Inc.
[5] Culick F.E.C., Unsteady Motions in Combustion Chambers fo r Propulsion Systems. 2006,
RTO/NATO.
[6] Lieuwen T.C. and Yang V., eds. Combustion Instabilities in Gas Turbine Engines:
Operational Experience, Fundamental Mechanisms, and Modeling. Progress in
Astronautics and Aeronautics Vol. 210. 2005, AIAA.
6 Introduction
[7] Peters N„ Turbulent Combustion. 1st ed. 2000, Cambridge: Cambridge University Press.
[8] Lipatnikov A., Fundamentals o f Premixed Turbulent Combustion. 2012, CRC Press.
[9] Markstein G.H., Nonsteady Flame Propagation. 1964, Oxford: Pergamon Press.
[10] Bellows B.D., Bobba M.K., Seitzman J.M., and Lieuwen T„ Nonlinear flame transfer
function characteristics in a swirl-stabilized combustor. Journal o f Engineering fo r Gas
Turbines and Power, 2007, 129(4): pp. 954-961.
Overview of the Book
This section previews the structure and content of this book and provides suggestions
for how readers of different backgrounds can use it most effectively. The bulk of
Chapter 1 is dedicated to reviewing the basic equations to be used in this text. Then,
the remainder of the book is divided into three main sections: Chapters 2-6, 7-9, and
10-12. The first section. Chapters 2-6, discusses flow disturbances in combustors.
Chapter 2 details how different types of disturbances arise and propagate in inhomo
geneous, reacting combustor environments. By introducing the decomposition of flow
disturbances into acoustic, vortical, and entropy disturbances, this chapter sets the
stage for Chapters 3-6 which delve into the dynamics of disturbances in inhomo
geneous environments in more detail. Specifically, Chapters 3 and 4 focus on the
evolution of vortical disturbances in combustor environments. Chapter 3 provides a
general overview of hydrodynamic stability theory and details some general features
controlling the conditions under which flows are unstable. Chapter 4 then details
specific canonical flow configurations that are particularly relevant to combustor
environments, such as shear layers, wakes, and swirling jets. This chapter also
discusses effects of flow inhomogeneity and acoustic forcing effects on flow
instabilities.
Chapters 5 and 6 treat acoustic wave propagation in combustor environments.
Chapter 5 provides a general introduction to acoustic wave propagation, boundary
conditions, and natural acoustic modes. Chapter 6 then provides additional treatment
of the effects of heat release, mean flow, and complex geometries on sound waves.
This chapter also includes an extensive discussion of thermoacoustic instabilities.
The second section of the book. Chapters 7-9, incorporates reacting flow phenom
ena and kinetics. Chapter 7 introduces the surface dynamics of propagating surfaces (a
model for premixed flames), constant-property surfaces (nonpremixed flames), and
material surfaces (nondiffusive passive scalars), illustrating their similarities and
distinctives. It also discusses how flames influence the bulk flow held, but does not
treat internal flame processes explicitly. Rather, it focuses on the influence of a flame
on pressure, entropy, vorticity, and velocity fields. Chapter 8 then treats auto- and
forced ignition. Chapter 9 covers flames, first reviewing premixed and nonpremixed
fundamentals, then moving on to more complex topics such as flame stretch, flame
extinction, and edge flames.
The third section of the book. Chapters 10-12, treats transient (in addition to the
ignition processes discussed in Chapter 8) and time-harmonic combustor
Overview
This chapter presents the key equations for a multicomponent, chemically reacting
perfect gas which will be used in this text [1], These equations describe the thermo
dynamic relationships between state variables in a perfect gas, such as the interrela
tionship between pressure, density, and entropy. They also describe the physical laws
of conservation of mass, which relates the density and velocity, the momentum
equation, which relates the velocity and pressure, and the energy equation, which
relates the internal and kinetic energy of the flow to work and heat transfer to the fluid.
This chapter’s primary purpose is to compile in one place the key equations to be
used throughout the text, and assumes that the reader has some prior familiarity with
them. A number of references are provided to readers for further details and deriv
ations of these equations. It is not necessary to follow the derivations to understand the
subsequent chapters, although an understanding of the physics embodied in each
equation is critical. For these reasons, discussions of various terms in these expres
sions are included in this chapter.
We will use the following perfect gas equations of state:
P & hT
( 1. 1)
MW '
( 1.2)
YfMW
(1.4)
MWj
10 Basic Equations
h= Y . Yi cpj(T * )d T * + J 2 hL Yi = cp(T*)dT* + Y J hl i Y¡ ■
i= 1 ¿=1 ¿=1 ¿=1
fochem
(1.5)
The first and second terms on the right-hand side of this expression are the sensible
enthalpy, hsem, and chemical enthalpy, of the system. The internal energy and
enthalpy are related as follows:
Additionally, the “stagnation” or “total” enthalpy and internal energy, defined as the
enthalpy or internal energy of the flow when adiabatically brought to rest, are given by
The rest of the chapter gives an overview of the conservation of mass, momentum,
and energy equations, while also providing evolution equations for other important
quantities such as fluid dilatation, entropy, vorticity, and kinetic energy.
^ + W - ( p u ) = 0. ( 1.8)
D
( ) + u ■V ( ), (1.9)
Dt dt
it can be cast in the alternative form
1 Dp
T V • it — 0. ( 1.10)
p Dt
Physically, this equation states that if one follows a given (i.e„ fixed mass) packet of
fluid, the normalized time rate of change of its density is equal to the negative of the
local divergence of the velocity. The local velocity divergence is, itself, directly
proportional to the rate of change of volume of a fluid element; i.e„ its “dilatation
rate,” given by the symbol A = V • it. Moreover, the dilatation rate is also equal to
the instantaneous flux of fluid out of a differential volume element of space. This can
1.3 Momentum Equation 11
be seen by integrating the dilatation rate over a volume and utilizing Gauss’s diver
gence theorem:
MV u ■ n dA. ( 1.11)
The resulting surface integral equals the instantaneous volume flux of fluid through
the control surface.
du _ _ Vp V - r " -
—— |- u • V it = — + ----= + VTuF,-, ( 1.12)
dt P P ^¿=1
and ¡Li, pÀ are the first and second coefficients of viscosity, S is the Kronecker delta unit
tensor, and S is the symmetric “strain rate tensor,’’ given by
1 ■ r
5 V it + (V it ) (1.14)
2 .
The momentum equation is an expression of Newton’s second law, stating that the net
acceleration of a fixed fluid element (Du /D t) equals the force per unit mass exerted
on it. The force terms on the right-hand side denote surface forces due to pressure and
viscous stress, and body forces. We will write three useful rearrangements of the
momentum equation next.
First, we can write a vector equation for the flow vorticity Q. = V x u by taking the
curl of Eq. (1.12):
Vp V t
V + V x ( u • V u ) = —V V Vx » / x •
dt P
(1.15)
Expanding further.
1 I—|2
— (V x u ) + V x ( V ( —| u | ) + (Q x m)
V t (1.16)
Vp x Vp
- V x Vp V v •
P . i =1
12 Basic Equations
DÛ
(V • u ) Û +(Q •V) u + * VP + V x
~Dt P-
(1.17)
Section 1.6 discusses the terms in this equation.
We next write a scalar equation for the evolution of the fluid dilatation A by taking
the divergence of Eq. (1.12):
DA Vp-(V -r) ^
p ---- = —p VY7-
it : VY7-
it —V p H, ^ P - ^ P + V •(V •r ) ------------- - f p V v . f h F i ) .
~Dt p p u
(1.18)
The double dot product appearing in the first term on the right-hand side is expressed
in tensor notation as
diti ditj
p V it : V it = (1.19)
^ dxj dx¡ '
The last equation to be developed from the momentum equation relates to the
kinetic energy per unit mass, | it | /2. This is obtained by taking the dot product of the
velocity vector it with Eq. (1.12). This leads to the scalar equation
We will utilize all of these momentum equation variants in the following sections and
subsequent chapters.
8- ^ ^ + V - ( p Y i u i) = wi. (1.21)
at
Note the distinction between the mass-averaged fluid velocity, it, and the velocity of a
given species, m,. Their difference is the diffusion velocity, m, — it = u d j - This
species conservation equation states that the time rate of change of a fixed mass of a
given element of species i equals the rate of production or consumption by chemical
reaction, Wj.
It is typically more useful to replace the species velocities m, in this expression with
the bulk gas velocity, it, and diffusion velocity, itr>.i- This leads to
1.5 Energy Equation 13
The reader is referred to Refs. [1, 2] for discussion of the terms in this expression,
which describe the diffusion of mass by gradients in concentration, pressure, and
temperature, as well as a species-dependent body force term. Equation (1.23) leads to
the familiar Fickian diffusion expression below for a binary mixture when Soret,
pressure, and body force terms are neglected:
»/>.; = /V In (1.24)
This expression can be used to recast the species conservation equation as the
following unsteady convection-reaction-diffusion equation:
DYj
P Dt viy + V • ( p S ’^ Y i ) . (1.25)
The species equations can also be recast into “conserved scalar” equations that are
source free. The key idea is that although atoms may move from one compound to
another during chemical reactions, they themselves are “conserved” and not created or
destroyed. For example, one can define conserved scalars based on each particular
atom, such as the hydrogen atom, H, and combine the different H-containing species
equations (e.g„ H2, H20 ) to arrive at a source- and chemistry-free equation [2, 3], This
approach is particularly useful in a system with a single fuel and single oxidizer stream
with the same mass diffusivities, where the mixture fraction, Z, is commonly defined
as the mass fraction of material originating from the fuel jet. The governing equation
for the mixture fraction is given by (see Exercise 1.8):
8Z
p — + p u - V Z - V - { p £ < V Z ) = 0. (1.26)
This is a statement of the first law of thermodynamics - the time rate of change of
internal and kinetic energy per unit mass of a given fluid element (the left-hand side)
equals the rate of heat transfer, V • minus the rate of work out (the latter three terms
on the right-hand side). There are three work terms, relating to work on the fluid at the
surfaces by pressure forces, V -(pit), and viscous forces, V - ( u -r), and by body
forces, ^2f=l{u + u d j ) • The heat flux vector ^ is given by
N N
X jS x i
-kTVT + p ^ h i Y , i M d , í ^ E E ( U D ,i— l l D , j ) + ‘¡ R a d -
'=1 3=1
(1.28)
See Ref. [1] for a discussion of these terms, which describe heat transfer by conduc
tion, mass diffusion, multicomponent mass diffusion due to concentration gradients,
and radiation, respectively.
Alternatively, the energy equation can be written in terms of total enthalpy,
hT = eT + (p/p). by moving the pressure work term from the right-hand side of
Eq. (1.27) to the left:
D hj
-V . + - ^ + V - ( m - t) + p Y¡(u + ud,í ) •F¡. (1.29)
1~d T dt
Equations for the internal energy and enthalpy can be obtained by subtracting the
kinetic energy equation, Eq. (1.20), leading to
De
p— = -V - -p (V • m) + t : ( V i í ) + / ) ^ f ¡ í í D,¡T ¡. (1.30)
¿=1
r\ N
Dh
p — = —V • ~ ~ + L : iy Û ) + P ^ Y i Û D ,i - F i- (1.31)
F Dt ¿=1
It is useful to explicitly bring out the chemical component of the enthalpy. This is done
by first writing an evolution equation for the chemical enthalpy of the system using the
species conservation equation, Eq. (1.22), as follows:
Dhcfiem N J3Y 1 _N 1 / _N
(1.32)
Dt 1=1
r=l rP i=\
r=l r \ i=l
Dhcfiem
q -X 7 - p ^ / V0,T,' IID, I (1.33)
Dt ¿=1
where the chemical source term q is the volumetric heat release rate due to combus
tion, given by
1.5 Energy Equation 15
This chemical enthalpy equation, Eq. (1.33), is then subtracted from Eq. (1.31) to
yield the following equation for sensible enthalpy:
Dh.,
V-
Dt
The energy equation can also be written in terms of entropy, l From Eq. (1.2),
Substituting Eqs. (1.30) and (1.25) into the above expression yields
Di N _ n l
pT— = - V • ^ + i : (V a ) - p S ’iVYi •Ft - + V • (pS',VT,)).
¿=1 ¿=1 1
(1.37)
This equation shows how the entropy of a given mass of fluid is altered by mass,
momentum, and energy diffusion, as well as body forces and chemical reaction.
Finally, the energy equation is written in terms of temperature and pressure. This
can be done by starting from the equation for sensible enthalpy and, noting that
dhsens = cpdT, rewriting Eq. (1.35) as
Dh..., Dp
p --------------- ? - V . | + r : ( V w ) + V. (p ^ hQ
f i YiUDJ ) + p YjUDJ ■Fj.
H Dt Dt ¿=1 ¿=1
(1.38)
The left-hand side of Eq. (1.38) can be expressed as the substantial derivative of either
temperature or pressure. For example, writing changes in temperature as
dT = (1.39)
DT T Dp TDp M W DYj
Dt p Dt p Dt rE
¿=1 MW: Dt '
(1.40)
so that
Dh., Dp DT Dp
Dt Dt PCp Dt Dt
T (\D p M W ^W i . \ / U \ ^ V - i p /V 1*/ *
PCpT[ y p D t+ ' l< p j ^ M W i MWj
/ 1 DT y — 1 / _ iviw MW 4 4 w. MW 4 ^ V ip /V 1*/ !
= pcpT --------- 1------- V • i t --------y
F p l yT Dt y l p ^ MWi - E - MWi
(1.41)
16 Basic Equations
ID T PCVT n nz MW¡
(y - 1)V- u = < N
T~Dt
b— —( —V- n +r_: (V m) + V- \ +p'ÿ^YiuDi-Fi
(1.42)
or for pressure as
q _+ » + I y - y - G oSW Y i]
1 Dp pCpT n n MW;
+ V- u — <
ypDt
+ p c j í — V- ^ + p : (V h) + V- ¡Uo j j + p ' ^ ^ Y i uDj-F ,
(1.43)
where n describes the time rate of change of the number of moles of the gas and is
defined as
Wi p
E MWj
------, n =
MW
. (1.44)
Note that
1 D ___ n
- = = — MW (1.45)
M W Dt n'
so that n = 0 if the average molecular weight is constant.
Neglecting molecular transport and body forces, Eq. (1.43) is
1 Dp _ _ à ñ
---------h YJ - it = --------1--- . (1.46)
yp Dt PcpT n
The two terms on the right-hand side of this expression are source terms that describe
volume production due to chemical reactions - the first because of unsteady heat
release and the second due to changes in the number of moles of the gas. To illustrate
the relative magnitudes of these two source terms, q and n, consider the ratio of
product to reactant gas volume, assuming constant pressure combustion:
where the superscripts u and b denote the reactant (unburned) and product (burned)
values. When fuels are burned in air, the n term is small relative to the q term, since
the reactive species are strongly diluted in inert nitrogen. However, there are applica
tions, most notably oxycombustion, where the molecular weight change between
1.6 Vorticity and Circulation Equations 17
This section discusses the various terms in the vorticity equation, reproduced below:
DO. A Vp x V« _ V t
~LH
{Q. ■V ) it £2 (V • i t ) ---------^------ f V EV
(1.49)
The left-hand side of this equation physically describes the time rate of change of
vorticity of a fixed fluid element. The right-hand side describes vorticity source or sink
terms. The first term, (Q ■V) it, is the vortex stretching and bending term. This term is
intrinsically three-dimensional and, therefore, is identically zero in a two-dimensional
flow. There are two processes described in this term, as illustrated in Figure 1.1. The
first process is the increase or decrease of vorticity by vortex stretching or contraction,
respectively. For example, consider a vortex tube oriented in the axial direction, Q v. If
this axial flow is accelerating, 8itx/e x > 0, then the tube is stretched, causing an
increase in axial vorticity, i.e., DQ.X/D t = Q.xdux/dx > 0.
The second phenomenon is the bending of a vortex tube originally inclined in
another dimension by the local flow. For example, a vortex tube that is initially
0 .0 .— 0 0 0 0 -
Flow acceleration
Shearing flow -
— - o
— o
— o
— o
Figure 1.1 Illustration o f stretching (top) and bending (bottom) effects on a vortex tube.
Basic Equations
Figure 1.2 Illustration o f processes by which a misaligned pressure and density gradient lead to
torque on a fluid element, and thus vorticity production.
oriented vertically, i.e., O ,, is rotated toward the x-axis in an axially shearing flow
cu j c y > 0. This rotation of the vortex tube causes it to be partially oriented in the
axial direction, inducing a Q v component, i.e., DQ.X/D t = i l xcu y/cy.
ReUtming to the full vorticity equation, the second term, (V • it) Q, describes flow
dilatation impacts on vorticity. It is only nonzero in compressible flows. Positive
dilatation, i.e., expansion in cross-sectional area of a vortex Uibe, leads to a reduction
in vorticity. This effect is analogous to a spinning skater who extends their arms
outward, leading to a slowing in angular velocity, or vice versa. This term has
important damping influences on vorticity as it propagates through the flame.
The third term, V;T V/I, describes vorticity production via the baroclinic mechanism,
which occurs when the density and pressure gradients are misaligned. This term is
identically zero in fluids where p = pip), e.g., in isentropic flows of a perfect gas. The
torque induced on the flow by a misaligned pressure and density gradient can be
understood from Figure 1.2. Consider the shaded control volume in a field with a
spatially varying pressure and density. The pressure gradient induces a net force on the
control volume, which acts through the center of pressure, shown in the figure.
Because of the density gradient, the center of mass is displaced from the center of
pressure. If the pressure and density gradients are not aligned (i.e., Vp x Vp / 0),
then the force acting at the center of pressure induces a torque on the fluid element
about this center of mass, creating vorticity.
Pressure only enters the vorticity equation through this baroclinic term. This
reflects the fact that only when the pressure and density gradients are misaligned
can the normal pressure forces exert a torque on the flow.
It is also important to note that vorticity can only be generated at no-slip boundar
ies, by baroclinic torque, or through body forces. All of the other terms in Eq. (1.49)
describe the amplification, stretching, bending, or diffusion of vorticity that already
exists in the flow. Thus, large-scale vortical structures which play such a key role in
Chapters 3 and 4 do not arise “from nothing” in flows without vorticity sources, but
from the complex reorganization of vorticity that enters the flow from boundary layers
1.6 Vorticity and Circulation Equations 19
(e.g., in jets, wakes, etc.). An enormous range of possible flow structures can arise
from this vorticity, depending on the characteristics of the specific flow in which
it arises.
More insights into sources of new vorticity, relative to reorganization/amplihcation
of existing vorticity, can be obtained by looking at the circulation. The circulation, T,
is defined as the surface integral of the vorticity or, equivalently, as the line integral of
the tangential velocity along the bounding surface, i.e..
T Q -dA= o •d t (1.50)
Si
DT 1 V •r \ - N
— (Vp x Vp) -dA + V ----= •dA +
~Dt hP- S i V x (Y iF i)-d A
(1.51)
Comparing this equation and the vorticity equation shows that, while stretching or
dilatation act as sources/sinks of vorticity, they do not appear as sources of circulation.
Baroclinic torque, diffusion, and nonconservative body forces are circulation sources/
sinks.
In addition, the manner in which vorticity induces fluid motion can be inferred from
Eq. (1.50). To illustrate. Figure 1.3 illustrates a two-dimensional region where vorti
city is confined to a region bounded by .S). Outside of Si, the flow is irrotational.
Equation (1.50) can then be manipulated to yield the relation
fld A ( ) it ■d t . (1.52)
Ji J2
In cases where the outer surface is much farther from the vorticity-containing region,
the value of ug along the circular bounding surface 2 will become progressively more
uniform, so that the line integral can be evaluated to yield
«9
Figure 1.4 M ie scattering image o f a nonreacting, J — 15 jet in cross flow. Dark regions
indicated by arrows are regions o f intense vorticity where seed particles are centrifuged out.
Image courtesy o f Vedanth Nair.
1
ue Q.rdA. (1.53)
2 nr
This equation shows that the induced velocity, ug, is proportional to the integral of the
vorticity. This induced motion has important practical consequences on the stability of
flows with multiple vorticity-containing regions, such as in wake flows where two
vortex sheets are convected from the bluff body as the boundary layer separates. These
vortex sheets mutually induce motion on each other, a fundamental driver of the von
Karman vortex street behind bluff bodies, as further discussed in Section 3.2.
This relation can also be useful for inferring average vorticity values from velocity
measurements around a bounding surface. For example, in particle image velocimetry
(PIV) measurements, there are often no seed particles in the vortex cores due to the
intense azimuthal velocities, such as shown in Figure 1.4. With no seed particles
present, the local velocity, velocity gradient, or vorticity cannot be measured in the
core region where the arrows point. However, by integrating the tangential velocity
along a bounding surface where sufficient seed is present, the average vorticity can be
determined from Eq. (1.52).
1.7 Nomenclature
This section details the nomenclature used in the text. Maintaining a consistent
nomenclature across the whole text is challenging given the different uses of common
1.7 Nomenclature 21
Ka Karlovitz number
L Length scale
Lp Flame length
¿11 Integral flow length scale
Lf] Kolmogorov flow length scale
Le Lewis number
ıh Mass flow rate
m". Mass flux; burning mass flux of flame
M Mach number
m Mode number
Ma Markstein number
MW¡, MW Molecular weight
ti Unsteady heat release gain factor
n Unit normal vector
n Molar production rate
n Number of moles, normal direction, axial acoustic mode number
P Pressure
Pr Prandtl number
î - ÎRad Heat flux
4 Chemical heat release rate per unit volume or flamelet
surface area
Q Heat release per unit mass of fuel reacted
Q Spatially integrated heat release rate
r Radial location
'fc o r r Correlation function or correlation coefficient
R Acoustic wave reflection coefficient
Coherence
Re, Reo, Rex, Re,-, Reynolds number
Real(X) Real part of X
Gas constant, universal gas constant
-áf -él Entropy per unit mass; entropy of species i per unit mass
sc, sd, s", sb Flame speed, speed of isosurface with respect to flow (sd)
S Strain rate tensor, see Eq. (1.14)
Sa Sankaran parameter, 1V |
SR Spin ratio, defined in Eq. (5.120)
St Strouhal number, fL /u
Std S t/ c o s 2 (<9), used in Chapter 12
Sti Strouhal number based on shear layer thickness, fr)/ux
Stp Strouhal number based on internal flame processes,fS p /s " ’°
Styr Strouhal number based on burner width, f W u / u x, used for
nonpremixed flame discussions in Chapter 11
Sy, Sm Swirl number
t Unit tangential vector
T Temperature
1.7 Nomenclature 23
3 Period of oscillations (1 / / )
T Acoustic wave transmission coefficient
Ux , U/y , U/£ ,
Ur , Ug. Un" . Un Scalar gas velocity component
—» -> -> ->
U, U ç , U i, Uj)^
~?b
U , u Fluid velocity vector
uc, UC,F Phase speed of velocity; flame wrinkle disturbance
V Volume
yF Velocity of flame surface
viV Mass-based production rate of species i by chemical reactions
w Duct height
W(y) Flame area distribution weighting factor used in Chapter 12
Xi Mole fraction of species i
Yi Mass fraction of species i
Y,n Bessel function of the second kind
z Acoustic impedance
Z Mixture fraction
Ze Zeldovich number
A Dilatation rate, V • u
P First coefficient of viscosity
Pi Chemical potential of species i
Pi Second coefficient of viscosity
Vv Dynamic viscosity, ¡i/p
e Angle or phase
Oj, Or* Onom Incident, refraction, and nominal angle
Opq Phase difference between pressure and heat release
P Density
Oj Temperature ratio
Op Density ratio
T Time scale
L Separation time
ign Autoignition time
T_ Shear stress
OJ Angular frequency
ÓJ Molar-based production rate of species i by chemical reactions
ñ Vorticity
s Interface position, usually referring to flame
v Stream function; spatial mode shape
C Decay, damping coefficient
1.7.3 Subscripts
0,1,2, ... Perturbations
A Reference position or value
B Reference position or value
c Flame speed: consumption-based definition, such as, ,v"; flow
velocity: disturbance convection velocity, uc
cun’ Curvature-induced flame stretch
D Displacement-based velocity definition, such as
displacement speed
>1 Associated with Kolmogorov scales
A Acoustic mode
i Imaginary
lean Lean mixture
MR Most reacting
n Normal component
NL Nonlinear part
P Potential mode
9 Azimuthal vector component (polar coordinates)
r Real part, or radial vector component (for polar coordinates)
i Entropy mode
Exercises 25
1.7.4 Superscripts
o Unstretched value of flame quantity
b Value of quantity at flame sheet on burned side
II Value of quantity at flame sheet on unburned side
Exercises
(Solutions are available at Cambridge.org/9781108841313)
1.1. A flow is isentropic if the entropy of a given mass of fluid does not change, i.e.,
D j/D t = 0. Work out the conditions that are required for this to be true.
1.2. Can a flow be isentropic, i.e., D j/D t = 0, and still have entropy fluctuations?
1.3. Work out the conditions for the stagnation enthalpy of a given mass of fluid to
remain constant, i.e., D h j/D t = 0.
1.4. The preceding problems worked out the conditions for the stagnation
enthalpy and entropy to remain constant along a pathline. Crocco’s equation
[4, 5] is a useful expression providing insight into the relationship between
flow and thermodynamic properties across streamlines (which is the same as
a pathline only in a steady flow). Moreover, it provides an enlightening
reformulation of the momentum equation into a form illustrating the
relationship between vorticity, stagnation enthalpy, and entropy. Starting
with Eq. (1.12), neglect body forces and molecular transport, and substitute
in the state relation
l1! dYi
dh TJ4 + J± +
p
E
1=1 W i1
M
(1.60)
26 Basic Equations
References
A key focus of this text is to relate the manner in which fluctuations in flow or
thermodynamic variables propagate and interact in combustion systems. In this
chapter, we demonstrate that combustor disturbances can be decomposed into three
canonical types of fluctuations - acoustic, entropy, and vorticity disturbances. This
decomposition is highly illustrative in understanding the spatial/temporal dynamics of
combustor disturbances [1], For example, the velocity held can be decomposed into
acoustic fluctuations, which propagate at the speed of sound with respect to the how,
and vorticity fluctuations, which are advected by the how. This decomposition is
important because, as shown in Chapters 11 and 12, two velocity disturbances of the
same magnitude can lead to very different influences on the hame, depending on their
phase speeds and space-time correlation. Section 2.9 further emphasizes how this
decomposition provides insight into behavior measured in a harmonically oscillating
how held.
This chapter is organized in the following manner. Section 2.1 introduces the basic
approach for analyzing disturbances, and illustrates the formal process of perturbation
expansions used throughout the text. Section 2.2 then considers small-amplitude
disturbance propagation in homogeneous hows. This limit is helpful for understanding
key aspects of the problem, as the disturbance modes do not interact and are not
excited. Section 2.3 closely follows this material by showing how these disturbance
modes are excited and how they interact with each other. Section 2.4 then considers
the energy density and energy hux associated with these disturbances.
Section 2.5 moves from specihc analysis of gas dynamic disturbance to a more
general overview of linear and nonlinear stability concepts. This and subsequent
sections pull together a number of results associated with linear and nonlinear
systems, the effects of forcing on self-excited systems, and effects of interactions
between multiple self-excited oscillators. These results have been selected from more
general nonlinear dynamics treatments for their relevance to important unsteady
combustor phenomena, which will be referred to throughout the text.
We will denote the “base” or “nominal” flow as the value of the quantity in the
absence of perturbations. This quantity is given the subscript 0, so in this case
Po = >j \. It is often useful to expand the quantity about this base state in a Taylor
series:
P = Po + Pi + P i H-----• (2-2)
where, by assumption, p 1 <iCp(j, Pi <C p \ , and so forth.1 Thus, for this problem we can
determine the form of these various orders of approximations by expanding Eq. (2.1)
to yield
p'(t) = p ( t ) - p . ( 2 .6)
In many problems, flow fluctuations can be further decomposed into random and
deterministic or quasi-deterministic fluctuations. For example, coherent flow structures
often consist of quasi-deterministic flow features, superposed upon a background of
fine-scale turbulence. Similarly, pressure fluctuations in thermoacoustically unstable
combustors are superpositions of broadband turbulent flow and combustion noise, and
near perfect tones at one or more frequencies. A “triple decomposition” can be used to
differentiate these different types of fluctuations, as described in Section 2.8.
Note that the time-averaged value is not equal to its base or nominal value; i.e.,
p / p(). Rather, p and p 0 are only equivalent to brst order in perturbation amplitude.
For instance, in this example problem,
PM sy
)’()’- 1 ) 0 - 2 )
sin (cot) fiQ Q - 1) cos (2cot)
Po 4 ( 2 .8)
fi3}’0 — i ) 0 —2) .
sin (3cot).
24
In general, the base value, ( )0 , of a quantity is not experimentally accessible, but its
time average is (similarly, the ( )j, ( )2, ... perturbation terms are not experimentally
accessible in general). For example, in a turbulent flow, the base flow consists of the
laminar flow that would exist at those same conditions. However, since the flow is
unstable, it is impossible to observe what this base flow would look like. Nonetheless,
expanding about a base flow is useful for approximate analytical techniques to
determine whether such a flow is linearly stable. Moreover, analysts often use the
formulae developed from such expansions and apply them to turbulent flows by
replacing the base flow values by the time-averaged ones. We will discuss this
specibcally in Sections 3.5 and 11.5.
With this distinction in mind, theoretical expressions developed from perturbation
expansions will be presented in this text as p 0, , p 2, and so on. In contrast,
measurements or computations are reported in terms of p and //.
ORIGINAL CAST