You are on page 1of 41

Progress in Energy and Combustion Science 35 (2009) 57–97

Contents lists available at ScienceDirect

Progress in Energy and Combustion Science


journal homepage: www.elsevier.com/locate/pecs

Ignition of turbulent non-premixed flames


Epaminondas Mastorakos*
Hopkinson Laboratory, Department of Engineering, University of Cambridge, Cambridge CB2 1PZ, UK

a r t i c l e i n f o a b s t r a c t

Article history: The initiation of turbulent non-premixed combustion of gaseous fuels through autoignition and through
Received 10 March 2008 spark ignition is reviewed, motivated by the increasing relevance of these phenomena for new
Accepted 21 July 2008 combustion technologies. The fundamentals of the associated turbulent-chemistry interactions are
Available online 16 September 2008
emphasized. Background information from corresponding laminar flow problems, relevant turbulent
combustion modelling approaches, and the ignition of turbulent sprays are included. For both auto-
Keywords:
ignition and spark ignition, examination of the reaction zones in mixture fraction space is revealing. We
Turbulent non-premixed
review experimental and numerical data on the stochastic nature of the emergence of autoignition
Ignition
Spark kernels and of the creation of kernels and subsequent flame establishment following spark ignition,
autoignition aiming to reveal the particular facet of the turbulence causing the stochasticity. In contrast to fully-
flame propagation fledged turbulent combustion where the effects of turbulence on the reaction are reasonably well-
DNS established, at least qualitatively, here the turbulence can cause trends that are not straightforward.
Autoignition occurs usually away from stoichiometry at a ‘‘most reactive mixture fraction’’, which can be
approximately determined from homogeneous or laminar flow autoignition calculations, and at locations
in the turbulent flow with low scalar dissipation. Such locations may be the cores of vortices. Once
autoignition has occurred at a time that is mostly affected by the history of the conditional scalar
dissipation, the relative magnitudes of convection, diffusion and reaction can affect the stabilisation
height of flames in sprays or jets. Modelling efforts based on the Conditional Moment Closure, advanced
flamelet approaches, and the transported PDF method seem suitable for capturing many, but not yet all,
of the trends observed in DNS or experiment. Further experiments and DNS of realistic fuels and at
conditions demonstrating chemical complexities must be performed to examine more fully the effects of
scalar dissipation and its fluctuations on pre-ignition reaction zones. The statistics of the first appearance
of autoignition in transient problems and its connection with the mixing field must also be studied.
Ignition from a localised spark has a stochastic character that depends on the mixture fraction sampled at
the spark location and duration and the local scalar dissipation. The success or not of the subsequent
flame depends on the development of turbulent edge or stratified flames. Only preliminary data exist on
the propagation speed of such flames and on their quenching. A lot remains to be done on turbulent edge
flame propagation in unreacted and partially-reacted mixtures with inhomogeneities, turbulent flame
propagation in non-uniformly dispersed droplet mists, and the transient stabilisation process of recir-
culating flames. The nature of the flame generation process at very short timescales, i.e. before any
appreciable propagation, by sparking in inhomogeneous mixtures needs also to be examined.
The development of high repetition rate diagnostics, for single- and two-phase flows, and the devel-
opment of modelling approaches capturing both premixed and non-premixed reaction zones in gaseous
and spray combustion are necessary.
Ó 2008 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
1.1. Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
1.2. Basics of autoignition and spark ignition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
1.3. Classification, applications and scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

* Tel.: þ44 1223 332 690; fax: þ44 1223 332 662.
E-mail address: em257@eng.cam.ac.uk

0360-1285/$ – see front matter Ó 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.pecs.2008.07.002
58 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

1.4. Additional literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62


2. Autoignition of turbulent non-premixed flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
2.1. Autoignition of laminar non-premixed flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
2.2. Ignition kernel location and structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
2.2.1. Most reactive mixture fraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
2.2.2. Low scalar dissipation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
2.3. Ignition times . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
2.4. Flame growth following autoignition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
2.5. Flame stabilisation mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
2.6. Modelling approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
2.6.1. Mean reaction rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
2.6.2. Transported PDF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
2.6.3. Laminar flamelet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
2.6.4. Conditional Moment Closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
2.6.5. Other . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
2.7. Spray autoignition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3. Spark ignition of turbulent non-premixed flames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3.1. Insights from premixed systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3.1.1. Ignition sources and the minimum ignition energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3.1.2. Spark ignition of turbulent premixed flows . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
3.2. Spark ignition of laminar non-premixed flames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
3.3. Ignition probability: flammability factor, Pker and Pign . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
3.3.1. Jets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
3.3.2. Counterflow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
3.3.3. Recirculating flames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
3.4. Flame growth following forced ignition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.4.1. Flame classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.4.2. Turbulent stratified-charge premixed flames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.4.3. Turbulent edge flames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.4.4. Flame stabilisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
3.5. Modelling approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
3.6. Spark ignition of sprays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
3.6.1. Homogeneous dispersions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
3.6.2. Turbulent sprays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
4. Conclusions and future research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
4.1. Autoignition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
4.2. Spark ignition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.3. Spray ignition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

1. Introduction operating the combustor lean. We see therefore that there exist
significant practical interest in autoignition and spark ignition. In
1.1. Background these applications, as in most other realistic combustion devices,
the fuel and air are flowing in a turbulent manner. Hence, we need
The autoignition (or spontaneous ignition or self-ignition) and the to understand and, eventually be in a position to predict, auto-
spark ignition (or forced ignition or assisted ignition) of a flammable ignition and spark ignition in turbulent flowing mixtures. This
mixture are fundamental problems in combustion science. They paper reviews the state of our knowledge on this topic, with
deal with the transition from an unreacted (in the case of spark particular emphasis on the initiation of turbulent non-premixed
ignition) or slowly reacting (in the case of autoignition) state to gaseous fuel combustion.
a fully or vigorous burning state that corresponds to combustion at
high temperature. For uniform mixtures with no flow, both prob- 1.2. Basics of autoignition and spark ignition
lems are usually treated in typical combustion textbooks (e.g. Refs.
[1–4]). In this paper, the term spark ignition is used interchangeably
Often in turbulent combustion research, attention is given only with (the perhaps more accurate for our aims) forced ignition. Both
to the fully-burning state, for example in order to discuss heat denote that an external source of heat or species is present to
release rates, pollutant emission or the possibility of flame induce the transition from the frozen to the burning state. We do
extinction. However, many technological applications focus not pay much attention to the exact nature of this source. The term
specifically on the transition to this state. The spark-ignition engine, autoignition is to be contrasted to forced ignition in that no external
the relight of an aviation gas turbine, and accidental spark ignitions source is needed to reach fully-fledged combustion. In either case,
of flammable gas releases are examples of forced ignition. The the effect of an externally-imposed pressure rise is to be explicitly
diesel engine, the spontaneous ignition of flammable releases, and distinguished. Glassman [4] discusses autoignition and forced
the scramjet are examples where autoignition is necessary to ignition succinctly in his Chapter 7 and the reader could benefit
initiate combustion. Recently the Lean-Premix-Prevapourised (LPP) from a reminder of the fundamentals.
gas turbine engine, where autoignition is to be avoided, is proposed It is instructive to discuss both phenomena through the various
as a way to achieve low NOx emissions due to the possibility of terms in the energy equation (for quantitative results, the full
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 59

system of equations including species conservation equations must mixture by sudden insertion of a layer of burnt gas. The burnt gas
be studied). In terms of the temperature, conservation of energy may be the result of having already introduced a finite qsp in the
can be expressed as layer, but the way the burnt gas layer has been generated is mostly
irrelevant for the ensuing flame propagation process. Here, T0 is so
  X N
vðrTÞ vðruk TÞ vp v vT low that we consider the chemical source term to be virtually zero
cp;m þ cp;m ¼ þ l  ha wa at the unreacted fluid. This configuration, by considering a second
vt vxk vt vxk vxk a¼1
spatial dimension and flow in this direction rather than time, can
N
vT X vYa model pilot ignition of a flammable mixture. To solve this problem,
þ rD cp;a þ qsp  qloss
vxk a ¼ 1 vxk we use a one-dimensional form of the energy equation with the
unsteady, diffusion, and chemical source terms only [1]. It is then
(1)
found that, for successful premixed flame propagation, the spark
The main quantities appearing in this equation are the velocity ui, must deposit enough energy to raise to the adiabatic flame
the mixture density r, the pressure p, the mass fraction of species temperature an area of size of the same order as the thickness of
aYa with D its diffusion coefficient (taken here equal for all species a laminar premixed flame at the conditions (temperature, pressure,
to facilitate presentation), the chemical source term wa, the stoichiometry) of the unreacted gas. Similar results exist for
conductivity l, the temperature T, and the absolute enthalpy ha, i.e. cylindrical and spherical sparks. (Quantitatively, the minimum
the sensible enthalpy plus the standard enthalpy of formation (e.g. spark size must be larger than the laminar flame thickness and is
Ref. [5], Chp. 1). The mixture specific heat capacity is given by best understood through the concept of quenching diameter [3,4].)
PN
cp;m ¼ a¼1 Ya cp;a , where N is the total number of species Experiments with forced ignition usually give the minimum ignition
comprising the mixture and the mixture enthalpy by energy (MIE), which is the minimum energy deposited in a partic-
PN
h ¼ a¼1 Ya ha . We have assumed a low speed flow so that viscous ular mixture necessary to initiate flames around the region of
heating and acoustic interactions are negligible. These may be energy deposition. Data for the MIE of various fuels are available
important for some problems, e.g. in detonations or in supersonic [4,6] (and such data have been extended to include droplets and
combustion, but such problems are not considered in this review. turbulence [7,8], as discussed in more detail later in this paper). In
In the case of forced ignition, the term qsp on the r.h.s. of Eq. (1) the context of Eq. (1), the spark will fail if it is too thin (i.e. qsp > 0 in
denotes the localised external source. Hence qsp will be finite only a very small region only) or not strong enough (i.e. qsp is too low).
in a small region in space and only for a short time; it is zero The magnitude of the energy source relative to the other terms can
otherwise. In the case of autoignition, qsp ¼ 0. The term qloss is therefore determine whether the jump from the unburnt to the
added to imply a volumetric heat (or species) loss. burnt state will happen in this fully-premixed situation.
The simplest problem of autoignition of a uniform stagnant fuel- In the steady laminar premixed flame, the convection, diffusion
air mixture starting from temperature T0 is solved by keeping only and chemical source terms balance. In a steady laminar non-pre-
the unsteady and the chemical source terms of Eq. (1). Due to the mixed flame, it is mostly the balance between diffusion and
increasing reaction rate with increasing temperature, even the very chemistry that determines the combustion. A canonical problem
weak reaction at T0 generates a little heat, which increases the for this case is the steady counterflow mixing layer between fuel
temperature further, thus providing a faster heat release rate. and air, where the diffusion and convection terms both scale with
Thermal runaway eventually occurs at an autoignition time that the strain rate and hence Eq. (1) reduces to an one-dimensional
depends on T0, the nature of the fuel, the reactants concentrations, ordinary differential equation, with the strain rate effectively
and the pressure.1 A variant of this problem includes volumetric summarising all the fluid mechanics and mixing of the problem. It
cooling (to mimic, for example, heat losses to a wall). The magni- is found [3,9] that a high strain rate can extinguish an already-
tude of qloss relative to the chemical source term can cause burning flame, as evidenced by the fact that burning solutions are
a fundamental change in the behaviour of the system. With found only if a suitably-defined Damköhler number is above
excessive heat loss, autoignition will not happen at all and the a critical value. The underlying concept of competition between
system will stabilise at a temperature only slightly higher than T0. chemistry and diffusion extends to autoignition: it is also found
With mild heat losses, the autoignition time will be finite, but that a high strain rate can preclude the autoignition of a fuel
longer than without heat losses. The analysis emphasizes that the flowing against hot air. This demonstrates that the balance between
reason behind the emergence of a transition is the very non-linear the convection/diffusion and the chemical source terms in the
dependence of the reaction rate on temperature due to the energy equation determine the transition from the frozen to the
Arrhenius expression and the high activation temperature of burning state. In an unsteady non-premixed laminar flow starting
combustion reactions. Because of the very sharp increase in the rate from T0 and unreacted conditions, if the chemical source term is not
of temperature rise at autoignition, the exact definition of the identically zero at this temperature, the transition to full combus-
autoignition point is usually not of any serious consequence. The tion will occur after an autoignition time that now depends on the
reader can consult combustion textbooks for details. mixing (i.e. the diffusion term) in addition to the chemistry [10].
The technological concept of autoignition temperature will not be These points will be discussed in some detail later in this paper.
used in this review. It is not a fundamental property of any material, The extent of our knowledge on the forced ignition of non-
as it simply denotes the T0 above which a given fuel-air mixture will premixed flames is very limited compared to what we know about
autoignite in a given closed vessel. The autoignition temperature is their autoignition. The existing work (reviewed in Section 3) shows
a useful practical device to rank the ease by which various that critical conditions concerning spark energy and strain rate
compounds autoignite, but cannot help us understand the details of exist also for the forced ignition of non-premixed combustion.
the autoignition process and how this is affected by the flow. It is not straightforward to extrapolate these insights to turbu-
The most common textbook problem of forced ignition is the lent flows. This is due to the multi-dimensional and unsteady
initiation of combustion in a premixed unburnt fuel-air uniform velocity fields, which cause temporal and spatial variations of the
mixture strength and the mass fraction and temperature gradients,
which in turn cause a wide range of relative magnitudes of the
1
In some cases, thermal runaway is not imperative for autoignition to happen, an
chemical source, convection and diffusion terms in the governing
accummulation of radicals at relatively constant temperature is sufficient. See, for equations. It is therefore not easy to comment on whether, when or
example, Ref. [2] for an exposition. where will the transitions from frozen to burning state take place.
60 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

We may wish to consider this review as an effort to establish the turbulent flows. With cold reactants, it has been used for spark-
turbulent counterparts of the simplified laminar flow problems ignition problems as well. This flow involves uniform strain and can
discussed above. demonstrate directly the effects of strain rate on the ignition
process and hence forms an important building block in our
1.3. Classification, applications and scope understanding of initiation of non-premixed combustion.
In Fig. 1, isolines of the stoichiometric mixture fraction, xst, are
It is useful to categorise various practical situations involving shown, together with the isolines corresponding to the lean and
turbulent non-premixed autoignition and spark ignition into rich static flammability limits (xlean and xrich, respectively). These
simpler problems that can be used to focus this review. This is limits are available for most fuels at atmospheric conditions.
attempted in this sub-section, with an introduction of a few However, they are not always available for the conditions corre-
important phenomena that we will be discussing in detail later. sponding to autoignition (e.g. the high pressure, high T2,0 of an LPP
Fig. 1a shows schematically a turbulent fuel jet into oxidizer (in turbine duct or a diesel engine) or to some practical spark-ignition
the following, the index 1 denotes the fuel-carrying fluid, while problems (e.g. at the low pressure, low temperature characteristic
index 2 the oxygen-carrying fluid; the index 0 denotes condition far of high altitude relight situations). A fourth curve, denoted as xMR, is
from the mixing region). In practice, the oxidizer stream could be meant to show the so-called most reacting mixture fraction – we
air or air mixed with hot combustion products or even air enriched don’t imply that xMR < xst necessarily. For cold fuel in hot oxidizer
with radicals. When U2,0 ¼ 0 and U1,0 is high, the flow pattern is (T2,0 > T1,0), autoignition happens somewhere along the xMR isoline,
akin to the usual turbulent jet. If T1,0 is low and T2,0 is high, the flow and in particular where the mixture fraction gradients, quantified
may correspond to gaseous fuel injection into a stagnant hot through the scalar dissipation N (N ¼ D(vx/vxi)2), are low [11]. The
environment, such as that in a natural gas direct-injection concept of the most reactive mixture fraction and the effect of N
compression-ignition engine. If the fuel is originally injected in and its fluctuations, first revealed through Direct Numerical
liquid form but the atomization and evaporation processes are very Simulations (DNS) and not yet fully examined by experiment, are
fast, then Fig. 1a may be considered conceptually as the pre-auto- very important for understanding turbulent non-premixed auto-
ignition part of diesel engine combustion. If U2,0 is high and T2,0 is ignition and are discussed in Section 2.
high (but perhaps not so high that autoignition happens quickly), Finally, Fig. 1d shows a typical situation of compression ignition
the flow mimics the premixing duct of an LPP gas turbine. in an engine with very early injection, which can be thought of as
Confinement and the co-flow turbulence are important in this case, a variant of a diesel engine usually denoted as Homogeneous
as they determine how quickly the fuel mixes with the oxidizer. Charge Compression Ignition (HCCI) (with the term ‘‘homoge-
This type of flow has been considered experimentally for various neous’’ very often not used in its strict sense). In the simplest case,
combinations of temperatures, velocities, fuel and oxidizer stream the fuel ‘‘blobs’’ (fluid 1) are being mixed with air (fluid 2). We
compositions, and for both confined and open flows and will be could envisage a situation with equal likelihood of ‘‘blob’’ distri-
reviewed in Section 2 of this paper. In a mixing layer (Fig. 1b), bution so that the fuel is homogeneous, in the mean, but with finite
contrary to the jet-type mixing flows, the limiting values of mixture fluctuations. Alternatively, we can say that Fig. 1d corresponds to
fraction are fixed (i.e. unmixed fluid is always available far from the a mixture fraction probability density function, P(h), that is uniform
layer), which allows easier interpretation of some autoignition in space and has finite width (h is the sample space variable of the
observations. This flow has been used extensively for autoignition mixture fraction x). Autoignition in this case is affected by the
simulations, but not yet for experiments. The opposed-jet flow pressure rise due to the compression. The difference between fluid
(Fig. 1c), with cold fuel impinging on hot air or combustion products 1 and fluid 2 could be only their temperature, i.e. Y1,0 ¼ Y2,0 but
has been used to study autoignition mostly for laminar, but also for T1,0 s T2,0. This problem does not fully qualify as non-premixed

T1,0
U1,0
ξrich
T1,0 ξrich
ξst
U1,0 ξMR
ξst
U2,0 U2,0
T2,0 T2,0 ξMR
ξlean ξlean

a b

U2,0
Y2,0
T2,0
T2,0
ξ
lean
ξMR

ξst Y1,0
ξ T1,0
rich

U1,0
T1,0

c d
Fig. 1. Schematics of canonical turbulent ignition problems included in this review. 1: fuel; 2: oxidizer. (a) Fuel jet in air co-flow; (b) Spatial mixing layer; (c) Opposed-jet flow; (d)
Volumetric compression with small-scale scalar fluctuations (species and/or temperature). For autoignition, T2,0 is usually high, while for spark ignition, T2,0 is low. The turbulence
may be pre-generated in both flows or it may be created by shear if U1,0 s U2,0. Combinations of these canonical problems may also exist in practice, for example (a) þ (d)
corresponds conceptually to a conventional diesel engine.
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 61

combustion, but its analysis can reveal some of the effects of along the xst isoline – and hence assumes characteristics more akin
turbulent fluctuations on autoignition. Dopazo and O’Brien [12] to non-premixed edge flames, whose behaviour will depend on the
were the first to consider such problems with the PDF method. local mixture fraction gradient. For low gradients, the flame may
In other HCCI concepts, fluid 1 in Fig. 1d could correspond to have three distinct fronts (triple flame: 3B) and for high gradients
a fuel-air mixture and fluid 2 to hot trapped combustion products, the fronts will have merged (edge flame: 3C). Such flames will be
so that initiation of combustion in fluid 1 may occur by a combi- denoted collectively as turbulent non-premixed edge flames and are
nation of autoignition (facilitated by the compressive heating) and reviewed in this paper in Section 3. Bilger et al. [19] have used the
diffusion from the hot fluid 2. Exactly which of the two phenomena term premixed/non-premixed for these flames, but we prefer the
is dominant has been a popular topic of discussion in the HCCI explicit inclusion of the word edge because tracking the physical
engine literature (a review is given in Ref. [13]) and the focus of flame edge has been used to measure the statistics of their propa-
fundamental studies [14–16]. The additional presence of diffusion gation. However, if the mixture fraction gradient is very low, no clear
from the recycled products in fluid 2 can help initiate combustion of edge may be discernible. The final outcome of the flame expansion
fluid 1 even under such lean conditions that its autoignition process will be a stable turbulent jet flame, attached to the nozzle or
wouldn’t normally occur. Another application where the distinction lifted, depending on the jet and co-flow velocity [20].
between autoignition and forced ignition (by a pilot) may be A special situation of flame propagation may arise if T2,0 is high
somewhat vague is in some scramjet concepts using continuous- and the flame kernel (i.e. kernel 1 in Fig. 2) has been generated
flow flameholders, such as plasma jets [17,18] or pilot flames. The through autoignition. In this case, the flame expansion process is
geometry would correspond to Fig. 1a or b with the oxidizer stream different, since the fluid ahead of the flame may be partially-reac-
at an intermediate temperature, but containing (perhaps in ted already, or even locally autoignited. This situation may corre-
pockets) extra radicals or hot products from the flameholder that is spond to the so-called premixed phase of conventional diesel engine
placed somewhere upstream. Flame initiation here should nor- combustion and some results on this are included in this paper
mally be classified as forced ignition of the non-premixed flow, (Section 2). We shouldn’t confuse this situation with forced ignition
since an external source is used. Strictly speaking, this is a three- in stratified-charge engines, where T1,0 and T2,0 are low enough so
stream non-premixed problem (fuel, air, pilot) and is not included that autoignition does not happen.
in this review. If perfect mixing between the pilot gases and the air Realistic industrial burners and gas turbine combustors virtually
has been achieved, the process could be considered as non-pre- always include strong swirl and recirculation. The recirculation is
mixed autoignition of the fuel stream in an enhanced (with radicals usually caused by the swirl and/or the action of a bluff body or
or hot products) air stream, a situation which is included here. a sudden expansion. A generic shape of the streamlines may be
Some features of the forced ignition in turbulent non-premixed similar to that indicated in Fig. 3, which also includes possible
flames are demonstrated in Fig. 2, which shows schematically the mixture fraction isolines. The mixture fraction pattern in these
flame expansion in a turbulent fuel jet in cold air (i.e. low T2,0) after geometries is extremely sensitive to the location of the fuel inlet
a successful spark. The most reactive mixture fraction is not a rele- and to the fuel velocity (magnitude and orientation). Qualitatively
vant quantity here because autoignition does not occur due to the different flame shapes may be observed [21], reflecting a very wide
low temperature and hence only the isolines of xst, xlean, and xrich are variety of mixture fraction distributions. The complex nature of the
indicated. Assume that a spark of high enough energy has been velocity and the mixture fraction fields in realistic burners intro-
deposited in a region comprising flammable mixture, for example duces further complexities to the flame expansion process, since in
between xst and xlean. Immediately following the spark, a small some locations the flow may assist while in others it may oppose
kernel has been generated, with an approximately spherical reaction the flame motion. Experimental observations show that even after
zone (1). The flame then expands (2). This flame is of a premixed expansion, the combustor is not always fully ignited, as the flame
nature, but spans a finite range of flammable equivalence ratios since may be quickly blown-off.
our flow is non-premixed; we can call this situation premixed flame Therefore, for the ignition of a burner, we must distinguish three
propagation in a stratified mixture, implying that the turbulent phases of the ignition process [7]. Phase 1: kernel generation involves
mixture fraction fluctuations fall within the flammability limits.
Bilger et al. [19] (their Chapter 4) have used this term for a situation ξlean
where fluid is either only fuel-lean or only fuel-rich (i.e. P(xst) ¼ 0)
ξst
and we will partly follow this convention here. Flame element 3A is
also of this character. As the flame continues to expand upstream, it
includes stoichiometric mixture fraction – it may actually move

3B ξrich
1

2
U1 ξst
ξrich
3A
3C
U2

ξlean

Fig. 2. Schematic of the kernel expansion process following localised ignition. For
autoignition, the flame expands in mixture that has already partially-reacted (and may
even have already autoignited at other locations too), while for spark ignition the flame
expands in unreacted fluid. 1: kernel immediately following ignition; 2: small turbu- Fig. 3. Schematic of a bluff-body stabilised non-premixed flame with radial fuel
lent stratified-charge premixed flame; 3A: turbulent stratified-charge premixed flame, injection just upstream of the bluff-body face, indicating mean streamline patterns and
3B: edge flame, low mixture fraction gradient; 3C: edge flame, high mixture fraction instantaneous xlean, xst and xrich isolines. The shape of the mixture fraction contours is
gradient. very sensitive to the fuel injection orientation and momentum.
62 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

the generation of a small flame around the spark. The second step, important to consider in isolation. We will also mention relevant
Phase 2: flame expansion, involves local propagation of the front results from laminar studies and from premixed systems that affect
relative to the unburnt fluid and turbulent dispersion of the already the behaviour of the turbulent non-premixed systems. We include
ignited fluid. The final step of Phase 3: flame stabilisation implies the a brief review of turbulent spray autoignition and spark ignition
long-term stability of the flame in the burner, with the combustion and we emphasize the many outstanding questions that need to be
having reached (statistically) steady-state conditions for the partic- addressed. A discussion on modelling approaches is included and
ular flow rates. All three phases must be successful for overall emphasis is placed on whether the physical phenomena included
combustor ignition and are separately discussed in this paper. Note or neglected in these can impact the accuracy and scope of the
that the phase of flame stabilisation is burner-specific. However, the predictions made.
generic geometry of a bluff-body stabilised flame (Fig. 3) may be used
to illustrate some relevant concepts through a discussion of some 1.4. Additional literature
recent results [22]. Phase 1 corresponds to the reaction zone outline
1 sketched in Fig. 2, while Phase 2 corresponds to outlines 2 and 3 in To the author’s knowledge, neither autoignition nor forced
Fig. 2. Phase 3 for the jet would be a normal jet non-premixed flame ignition in turbulent non-premixed flows have been the subject of
(implying usually a reaction zone pegged on the xst isoline). It will be a focused review so far. However, various relevant constituent
shown later that evidence exists than even for the jet and the phenomena have been reviewed in this journal and elsewhere. The
opposed-jet flows (Fig. 1c), which are probably the simplest non- reader would do well to consult these references. Aggarwal [8]
premixed configurations studied so far, Phase 3 may fail. discussed in detail the ignition of sprays (both autoignition and
Assume we perform a large number of individual spark trials in forced ignition), with emphasis on findings from laminar flows and
our turbulent non-premixed flow and we monitor the final from stagnant droplet-air mixtures. One of Aggarwal’s suggestions
outcome of each trial. Each time the flame is ignited, we extinguish for further research was to consider the autoignition and spark
it and we repeat the process. Due to the turbulent nature of the ignition of turbulent flows because very little work had been done
flow, the possibility exists that in some events of this ensemble the in this topic at that time. After a decade, enough knowledge has
spark may be located in, say, pure air, while in others the spark may been amassed to warrant a review.
lie in pure fuel. In the limit of an instantaneous and point spark, it The main chemical aspects of fully-premixed autoignition for
may be assumed that the statistics of the turbulent mixture fraction hydrogen and hydrocarbon fuels are presented to a reasonable
fluctuations at the spark location determine the probability of depth in textbooks [2,4] and extensively in various review papers
whether the spark samples flammable mixture or not. Hence the [26–28]. Law [2] includes the autoignition of strained counterflow
forced ignition of a turbulent non-premixed flow has a stochastic laminar mixing layers (his Chapter 8). The counterflow geometry
nature and the fluctuations of the mixture fraction are important; (Fig. 1c) will be discussed here as well, as it demonstrates one of the
we cannot treat it based only on the mean values. This is a central possible effects of the turbulence on the autoignition process. Some
concept, to the author’s knowledge first put forward by Birch et al. citations of analytical autoignition studies are given in Section 6 of
[23,24], who coined the term flammability factor, F, for the area Ref. [29]. Various aspects of droplet evaporation and autoignition
under the mixture fraction pdf between xlean and xrich. F is the are reviewed by Sazhin [30]. The interplay between autoignition
probability of finding nominally flammable fluid at a point and this and flame propagation or stabilisation in non-premixed systems is
probability, to a first approximation, can be used to discuss the also relevant to the so-called MILD combustion regime, whose
statistics of spark success. characteristics are reviewed thoroughly by Cavaliere and de Joan-
In reality, sparks may last for significant time compared to the non [31]. The effect of air vitiation on ignition in the context of
convection time of eddies past the spark location. In addition, the ground experiments of scramjet combustion has been discussed in
spark is of finite size, especially if we consider the very quick Ref. [32], including a review of some of the fundamentals of auto-
expansion following the initial energy deposition. These arguments ignition in laminar non-premixed systems, but with little emphasis
imply that the statistics of the mixture fraction sampled by the on the effects of turbulent mixing. Both autoignition and forced
spark depend on spark duration and size in addition to the turbu- ignition are relevant to supersonic combustion [33].
lence. Finally, the possibility that the kernel can be quenched from Lyons [20] reviewed the structure of turbulent lifted jet flames,
high local strain cannot be discounted, since we know that this may whose stabilisation point may be similar to the propagating edge
occur during forced ignition of turbulent premixed mixtures [25]. flames during flame expansion after spark ignition in inhomoge-
Therefore, the probability of initiating a kernel, Pker, is the result of neous mixtures (Fig. 2). Laminar edge flames have been reviewed
many factors and is not necessarily equal to F. In addition, in view of by Buckmaster [34] and in Section 3 of [29]. Here, we will add some
the discussion of flame propagation and stabilisation following recent results on turbulent edge flames. Information on electrical
a spark given in the previous paragraph, the probability of initiating sparks is provided in Ref. [35]. In more practical terms, correlations
a flame in the whole combustor, Pign, can be smaller than Pker and summarising spark ignition and flame stability in gas turbine
this distinction is emphasized in this paper. Insight into the combustors have been discussed by Mellor [36] and in the book by
behaviour of F, Pker and Pign and the differences between these Lefebvre [7]. Drake and Haworth [37] discuss various new low-
contribute to our understanding of forced ignition of non-premixed pollution internal combustion engine concepts such as HCCI and
flames and form a focal point of our discussions in Section 3. stratified-charge spark-ignition engines in the light of advanced
The flows shown in Figs. 1–3 may include droplets and hence diagnostics and available modelling approaches. The autoignition
the concepts discussed above are present also in spray autoignition of synthesis gas (mixtures of hydrogen and carbon monoxide
and spark ignition. In view of the importance of liquid fuels in derived from coal or biomass gasification), a fuel with significant
combustion applications, we include a brief discussion of ignition future potential, is discussed in recent reviews of practical appli-
phenomena in turbulent sprays. cations [38,39] and fundamentals [28].
The main objective of this paper is to consolidate existing Extensive literature exists discussing ignition in the context of
knowledge on the autoignition and spark ignition of turbulent safety, although in general the fundamental effects of turbulence on
non-premixed combustion of gaseous fuels, focusing on the flow the process are not discussed. A pressurized release of flammable
situations described in Figs. 1–3 and on the key phenomena gases may autoignite due to the shock wave associated with the
introduced in this sub-section. These phenomena can be under- sudden release [40] or it may create a flammable mixture that can
stood as sub-problems constituting the full problem and are hence be ignited by a spark and both possibilities must be considered for
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 63

safety issues. In view of a possible future extensive use of hydrogen cannot happen if the strain rate is above a certain value, Scrit, the
as fuel, a review of various mechanisms of ignition of hydrogen strain rate corresponding to this turning point. Due to the S-shaped
leaks has recently become available [41], highlighting that much nature of the curve, this critical value is lower than the critical
work remains to be done. Ignition of a solid flammable material strain rate for extinction of a flame. Hence, autoignition in the
under the action of radiation is discussed by Hirano [42], while opposed-jet geometry will not happen if the strain rate is higher
ignition of flames in dust clouds is reviewed by Proust [43]. than Scrit. Key results from the analysis on the critical autoignition
Turbulent combustion modelling has been reviewed by Vey- Damköhler number are described also in Ref. [49]. Explicit data are
nante and Vervisch [44], Bilger et al. [19], and a tutorial book is also available on the S-shaped response of maximum temperature or
available [45]. Pitsch discusses Large Eddy Simulations of turbulent a radical species concentration to scalar dissipation through
combustion in detail [46]. Peters’ monograph [47] includes a useful detailed chemistry simulations (for example Refs. [50,51]; see also
description of approaches for lifted flames, stratified-charge pre- Ref. [2] for a qualitative discussion.)
mixed flames, and for non-premixed autoignition. Here, we will The existence of a critical strain rate for autoignition in the
discuss the aspects of the existing models that are sufficient or may opposed-jet flow has been borne out by experiment for a wide range
need improvement for capturing turbulent autoignition, flame of fuels and operating conditions (dilution, temperature, pressure).
development following kernel generation from a spark, and overall In the experiments, the flow rates are fixed and the temperature of
burner stability. the oxidizer stream is gradually increased or the temperature is
fixed and the strain rate gradually decreased, until a flame appears.
2. Autoignition of turbulent non-premixed flows The result is the determination of a critical temperature Tcrit as
a function of strain rate, fuel dilution, fuel type, and pressure, in
In this section, we will discuss fundamental findings and addition to the nature of the oxidizer. Hydrogen [52], H2/CO
simulation methods for autoignition in turbulent non-premixed mixtures [53,54], methane pure [55] and hydrogen-enriched [56],
flows. Note that modelling of such flows has come earlier than these and various other single component hydrocarbons [49,57–60],
findings have become available, probably due to the practical methane-ethylene mixtures [61], dimethyl ether [62], and real fuels
significance of predicting the onset of combustion in diesel engines. such as aviation kerosenes [49,60] have been examined. It is found
In view of our aim to identify the strengths and weaknesses of the that virtually always Tcrit increases with S. An exception is hydrogen
models, we discuss these last, after we have introduced the key in a range of pressures and temperatures around the second
physical features revealed from DNS and experiment. explosion limit, where an insensitivity of Tcrit to strain is observed
The DNS work has, over time from the mid-90’s, extended from [52]. These experiments have usually been accompanied by detailed
2-D with simple chemistry to 3-D with complex chemistry. mechanism simulations, analyses of which have indicated the key
Experimental work is still extremely limited, with only a handful of elementary reactions participating in the autoignition of the various
experiments identified with refined enough results to reveal fuels at the particular conditions studied. Many of these mecha-
something about turbulence-autoignition interactions. There are nisms have been developed by reference to the correct prediction of
numerous experiments with sprays injected in a hot environment, Tcrit and of its dependence on S. Simulations additional to those in
as this configuration is very relevant to diesel engines, but these the above papers have been performed for hydrogen [63,64],
works in general do not focus on the turbulence effects on the heptane [65–67], and octane [67]. Law, in his book [2], reviews some
autoignition process. Modelling work covers most known turbulent of the chemistry of hydrogen at various pressures and how this
combustion approaches and is quite extensive. results in an autoignition behaviour that under some conditions can
In order to understand autoignition in turbulent non-premixed be understood as radical rather than thermal runaway [63]. This
flows, it is important to examine the autoignition of the corre- fundamental change in the chemical structure of the system, e.g.
sponding laminar flows and this is discussed in the next sub- switching from a chain branching to a thermal explosion mode, can
section. We then move on to a presentation of the main findings on cause the relative insensitivity of the response of the critical
autoignition location and time in turbulent flows. We close this temperature to the strain rate observed in the experiments.
section with a discussion of modelling approaches and a brief The effect of a periodic strain rate (or scalar dissipation) has also
review of turbulent spray autoignition. been numerically explored with detailed chemistry [50,51,68,69].
This problem reveals some interesting features. In particular, if the
2.1. Autoignition of laminar non-premixed flows time-averaged N is above Ncrit (Ncrit being the steady critical scalar
dissipation above which autoignition is not possible), autoignition
The most widely studied non-premixed configuration in terms may still occur depending on the amplitude and frequency of the
of autoignition is the laminar counterflow (Fig. 1c), where a fuel jet oscillation, which may allow an excursion into the low-N regime to
impinges upon a hotter oxidizer, with both streams shielded from last long enough to permit autoignition. Conversely, if the time-
the environment by nitrogen co-flows. This geometry, with rela- averaged N is below Ncrit, autoignition may still be precluded due to
tively low oxidizer temperature, has also been used very exten- long-lasting excursions above the critical value. These findings have
sively for laminar flame studies in terms of flame structure, some relevance to turbulent non-premixed autoignition, where the
response to strain rate, and eventual extinction.2 fluctuations of N are very important as we will discuss later.
In a very well-known paper by Liñán [9], it is shown that the Experimental data exist for autoignition of hydrogen [70] and
lower turning point of the S-shaped response of flame temperature methane [71] when the oxidizer stream is composed of the prod-
with the Damköhler number corresponds to autoignition, i.e. ucts of a lean flame, i.e. vitiated air. This mimics various combustion
a transition from a ‘‘frozen’’ to a burning state. This transition situations, such as the Moderate and Intense Low oxygen Dilution
(MILD) combustion [31]. The situation of a lean fuel-air mixture
impinging against vitiated air has also been studied [71,72], but in
2
In the laminar counterflow, the scalar dissipation is proportional to the strain principle, combustion initiation in this case should be considered as
rate and additionally depends on position across the layer. In this paper, we will use forced ignition of the mixture by the diffusive action of the
the two terms almost interchangeably, but when considering turbulent flows it is opposing hot stream. If the vitiated air is above a temperature close
preferable to use the scalar dissipation, since high fluid mechanical strain rates will
not necessarily imply high scalar dissipation rates, e.g. in a well-mixed turbulent
to the inner-layer temperature of the corresponding premixed
fluid. See, for example, Refs. [44,45,47,48] for a discussion of the importance of flame, there is no sudden autoignition and a reaction zone is always
scalar dissipation in turbulent combustion. evident for all fuel-air compositions and for all strain rates [71]. In
64 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

this case, the S-shaped curve becomes monotonic. This topic is fraction. Various complexities due to the chemical mechanism,
analysed in Refs. [31,72]. The particular effects of the various however, may be present especially in the intermediate species
species present in the vitiated air (such as the thermal effect of that appear before autoignition, such as H2O2 and HO2 for
major species or the chemical effects of trace species such as NO) hydrogen and a host of other species for hydrocarbons. Using
have been explored through numerical simulations and experiment detailed transport properties will also change these curves. A
[73–76], due to the increasing importance of this topic in HCCI common finding from examination of similar data reported for
engines [77] and in scramjet combustion [32,76]. various fuels suggests that the autoignition location in mixture
The unsteady strained mixing layer, with constant or variable fraction space is, in general, different from xst: depending on the
strain, has been simulated extensively and it is shown that fuel and the conditions, it can be at leaner (e.g. for hydrogen or
increasing the strain rate causes a retardation of ignition time methane, see Refs. [81,82]) or richer (e.g. for heptane, see
[51,68,69,78,79]. This is a major finding and is fully consistent with Ref. [83]). We will discuss this point in detail later in connection
the increasing Tcrit with S observed experimentally and numerically. with turbulent flow autoignition. The very small temperature rise
As in the steady counterflow problem, if S is above a critical value, before autoignition in Fig. 4 may be particular to hydrogen [63]; in
the autoignition time becomes infinity, i.e. we preclude auto- general a small, but observable temperature rise occurs. The
ignition due to too high strain rate. The transient flamelet equation reader is encouraged to always perform laminar transient diffu-
[44,47] for the temperature and mass fraction of species a, which is sion layer simulations, in either physical space or mixture fraction
also used for turbulent flames (discussed later), can be written as space, for any turbulent non-premixed autoignition problem
considered, and to plot results in terms of mixture fraction as in
vT v2 T Fig. 4.
r ¼ rNðxÞ 2 þ wT (2) Again for exposition purposes, a typical behaviour of the
vt vx response of the autoignition time with N0 is given in Fig. 5 from
a transient flamelet simulation of heptane [84] using unity Lewis
numbers. Autoignition time is best defined by examining the
vYa rNðxÞ v2 Ya temporal evolution of a quantity that increases significantly at the
r ¼ þ wa (3) burnt flame from its inert value, such as OH, which is closely related
vt Lea vx2
to the chemiluminescence emission from OH* that is used often in
Note that in Eqs. (2) and (3), N has been defined as N ¼ (l/ experiments. It is usually not very important what definition is
rcp,m)(vx/vxi)2, but if a unity Lewis number is assumed, the more used. The simulations show that sign increases with scalar dissi-
common definition of scalar dissipation based on the diffusivity of pation, at first mildly.3 As N0/N0,crit approaches unity, sign rapidly
the mixture fraction is recovered. For laminar layers, the mixture approaches infinity, i.e. no autoignition. From this graph we
fraction is a known function of cross-stream distance and hence conclude that N0,crit is about 150 1/s for the fuel and air tempera-
N(x) can be written as N(x) ¼ N0G(x), with N0 the maximum scalar tures, dilution and pressure used. The quantification of the critical
dissipation across the layer (at x ¼ 0.5) and G(x) a bell-shaped scalar dissipation through numerical simulations or experiments is
function peaking at unity at x ¼ 0.5 (e.g. Ref. [47]). Solution of Eqs. important for extrapolating results from laminar studies to turbu-
(2) and (3) starting from given initial conditions can be obtained lent non-premixed autoignition. Similar data have been produced
by numerical integration. Chemistry to any complexity can be for a range of fuels and conditions, with detailed chemistry and
used. transport (e.g. Refs. [51,65,68]), but more is necessary to cover the
For exposition purposes, a typical behaviour of the temperature range of practical engine operation.
and some species against the mixture fraction for various times is Practical experience shows that the numerical analysis here
shown in Fig. 4 for hydrogen using a detailed mechanism [80] and must be done very carefully because small errors in the calculation
a constant N0 ¼ 20 1/s, for unity Lewis number. The fuel and oxygen of radical and intermediate species can cause large differences in
are consumed little and the temperature at first increases by a very autoignition times and in the quantification of Ncrit. Autoignition, in
small amount, while some intermediate species (e.g. HO2; Fig. 4d) this respect, is much less ‘‘forgiving’’ to the numerical solver than
increase gradually during the induction time (this term will denote a flame. Fig. 5 also includes a comparison between a detailed and
here the time until autoignition occurs). The species related to a reduced mechanism derived from the detailed. The reduced
high-temperature reactions such as final products of the combus- mechanism was based on selecting steady-state species by the
tion and the temperature increase suddenly at a time associated Computational Singular Perturbation method [85]. Some differ-
with the autoignition time. For the example shown in Fig. 4, ences are evident, which highlights the more general observation
autoignition can be considered to happen at about 0.5 ms. Before that it is relatively difficult to construct reduced mechanisms for
this time, HO2 has been accumulating at mixture fractions between autoignition and that, for an equal level of accuracy compared to
0 and 0.2 with a peak below 0.1 (the stoichiometric mixture fraction the original detailed scheme, these reduced mechanisms tend to be
is xst ¼ 0.127). Autoignition is associated with a very rapid larger than for flames [86].
destruction of the pre-ignition species (HO2 in this case; see the In unsteady unstrained mixing layers between cold fuel and hot
curve at t ¼ 0.5 ms in Fig. 4d) and the major reactants H2 and O2, air, autoignition always eventually occurs, at least numerically
and a rapid generation of OH. The flamelet eventually assumes the [75,83,87,88] and in high activation energy asymptotics [10,89,90].
shape of the steady-state flame corresponding to the particular T1,0 This problem was analysed earlier by Dooley [91] in 1956 and
and T2,0. After autoignition occurs, the simulations show that a calculation for the autoignition time was made, but expressed in
reaction fronts propagate across the layer to consume the already- a cumbersome manner. That analysis included velocity shear. The
premixed fluid, giving rise to three (one lean, one rich, and one at
xst) reaction zones. The speed at which these fronts propagate
increases with N. If the scalar dissipation is higher than the critical 3
In these particular simulations that used unity Lewis number and with the
value, the flamelet becomes ‘‘frozen’’, with very small radical detailed mechanism, a very slight reduction in sign may also be observed at small
generation (Fig. 4f) and virtually no temperature increment over N0/N0,crit due to the diffusion of radicals from the high-temperature very lean
region, where these are first created, to richer mixture fractions, a process that is
the initial value.
accelerated with increasing N0 [84]. A monotonically-increasing sign with N0 is more
These results are qualitatively similar for autoignition of all common, however, with virtually no increase with N0 at very small N0/N0,crit usually
laminar non-premixed flows, when expressed in terms of mixture observed (e.g. Figs. 11 and 18, Ref. [65]).
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 65

a 0.10
t=0 b 0.40
t=0
t=0.1ms t=0.1ms
t=0.2ms t=0.2ms
0.08 t=0.3ms t=0.3ms
t=0.4ms 0.30 t=0.4ms

H2 mass fraction

O2 mass fraction
t=0.5ms t=0.5ms
t=0.6ms t=0.6ms
0.06 t=1.0ms t=1.0ms
t=2.0ms 0.20 t=2.0ms
t=4.0ms t=4.0ms
0.04

0.10
0.02

0.00 0.00
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Mixture fraction Mixture fraction

c 0.015 d 0.00020
t=0
t=0.1ms t=0
t=0.2ms t=0.1ms
OH mass fraction

t=0.3ms t=0.2ms
t=0.4ms

HO2 mass fraction


0.00015 t=0.3ms
0.010 t=0.5ms t=0.4ms
t=0.6ms t=0.5ms
t=1.0ms t=0.6ms
t=2.0ms t=1.0ms
t=4.0ms 0.00010 t=2.0ms
t=4.0ms
0.005
0.00005

0.000 0.00000
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Mixture fraction Mixture fraction

e 2400.0
t=0 f 5.0e−08
t=0.1ms t=2 ms
2000.0 t=0.2ms t=4 ms
t=0.3ms 4.0e−08 t=6 ms
HO2 mass fraction

t=0.4ms
Temperature (K)

t=8 ms
1600.0 t=0.5ms t=10 ms
t=0.6ms
t=1.0ms 3.0e−08
1200.0 t=2.0ms
t=4.0ms
2.0e−08
800.0

400.0 1.0e−08

0.0 0.0e+00
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Mixture fraction Mixture fraction

Fig. 4. Transient laminar flamelet calculations with unity Le for hydrogen-air constant-strain mixing layer with N0 ¼ 20 1/s (a–e) and N0 ¼ 70 1/s (f) between hot air and cold fuel.
For (f), the scalar dissipation is above the critical value. Yfu,0 ¼ 0.1, 1 bar, detailed mechanism from Ref. [80]. For the chosen dilution, the stoichiometric mixture fraction is 0.127.
Unpublished data, simulations performed with the code presented in Ref. [84].

asymptotic analysis of Liñán and Crespo [10] using one-step rates. Depending on the timescale of the chemistry relative to the
chemistry deserves special mention. In this work, the location of vortex, autoignition may occur at the centre of the vortex core or in
autoignition in mixture fraction space is derived, although not the strained regions outside the vortex [79]. More detailed analysis
given a large emphasis. We will revisit this point later. with detailed chemistry [92] shows that autoignition location
Differential diffusion can alter the calculated sign by a few correlates well with locations where the scalar dissipation is low (a
percentage points for hydrocarbon fuels [83], but by a factor of two result that has also been observed in DNS of turbulent autoignition
or three for hydrogen [81,88]. This is because the higher penetra- [11], to be discussed in detail later), demonstrating that in most
tion of the light H2 and H species into the hot air stream implies that laminar flows studied so far there is strong evidence that
further radical production is facilitated. The practical implication of autoignition is favoured with low scalar dissipation.
this observation is an earlier autoignition time with the addition of
hydrogen in hydrocarbon fuels (and a decreased Tcrit for strained 2.2. Ignition kernel location and structure
systems [56]) that is not only due to chemical reasons, but
additionally due to transport. However, further work is necessary to 2.2.1. Most reactive mixture fraction
identify fully the effects of non-unity Lewis numbers on auto- In Fig. 4, it was evident that the radical species and the
ignition of non-premixed flows. temperature increased preferentially at some mixture fractions,
We finally discuss another interesting problem from laminar which implies that autoignition will not happen at all mixture
flame simulations. A straight fuel-air interface, if disrupted by fractions simultaneously. Further, it seems that the x that first
a vortex, can produce mixtures that experience a range of strain ignites is separated from the stoichiometric.
66 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

3.0 with one-step chemistry in 2-D [94] and 3-D [95] with initial
mixture fraction distributions from a turbulent spectrum rather
detailed mechanism − VODPK
reduced mechanism − CHEMEQ2 than a straight fuel-air interface, simulations with hydrogen
reduced mechanism − VODPK detailed chemistry in 2-D [81,96,97], and with a four-step heptane
2.5 mechanism in 2-D [98] and 3-D [99]. In all these, the turbulence is
decaying. Problems with flow include the supersonic mixing layer
ignition delay [ms]

(in 2-D) [100], jets of hydrogen in slow co-flow of heated air (in
3-D) [101], and jets of hydrogen in fast co-flowing turbulent heated
2.0 air (in 2-D) [102]. An important difference of the detailed chemistry
results compared to the one-step results is that the location of the
peak of the conditionally-averaged heat release rate may shift by
a large amount towards the end of the induction time, and that this
1.5 shift depends on the scalar dissipation. However, this shift may or
may not be significant, depending on the fuel, the chemical
mechanism used and the conditions (e.g. T2,0 and T1,0; the effect of
pressure on xMR has not been studied explicitly yet). We can also
1.0 observe in Fig. 7 that the reaction zone has significant thickness in
1 10 100 1000
peak scalar dissipation rate N0 [1/s] mixture fraction space. This may have been due to the relatively
low activation energy used in the one-step chemical scheme used
Fig. 5. Calculations of autoignition time of heptane in an unsteady flamelet between for that simulation, but the fact that pre-ignition reaction zones are
hot air and cold fuel with unity Lewis numbers under constant scalar dissipation with relatively wide is borne out by detailed-chemistry DNS and laminar
reduced and detailed mechanisms. Reprinted from Ref. [84], with permission from flame simulations as well.
Elsevier. The potential effect on the results of the numerical solver is also evident. Is there a way that we can pre-compute xMR? Three separate
methods may be used, all giving results that are reasonably similar.
The same observations have been found in turbulent situations, First, a series of homogeneous reactor calculations can be per-
which constitutes our first insight into the autoignition of turbulent formed, with initial conditions of species and temperature corre-
non-premixed combustion, and has come virtually completely sponding to frozen mixing between the cold fuel and the hot air,
from DNS with various configurations, codes, chemistries and which can hence be fully parameterised by the mixture fraction x.
conditions. For most studies, T1,0 < T2,0, so autoignition is accom- Plotting the autoignition time sign from each of these vs. the cor-
plished as in the laminar canonical flows discussed in the previous responding x then results in a curve with a minimum. At very lean
sub-section. The first flow pattern studied was the temporal mixing mixtures, the temperature is high but there is little fuel. As x
layer (Fig. 1b). Initially, a straight interface separated the fuel from increases, the temperature decreases but the fuel concentration
the air and homogeneous isotropic turbulence was introduced. increases. The competition of these two effects results in an
Limited results were also obtained from slabs of cold fuel exposed optimum composition, giving the fastest autoignition time. The
to hot air from either side [11] and small ‘‘blobs’’ of fuel in air [93]. minimum autoignition time, called the reference autoignition time
The DNS code used for these simulations was NTMIX, developed at sref, can be used as a characteristic timescale of the autoignition of
CERFACS and IFP in France, and was a fully-compressible code with the particular non-premixed flow corresponding to no mixing. The
one-step chemistry simulating two-dimensional turbulent flows. mixture fraction at which this minimum occurs can be considered
No forcing was used and hence the turbulence was decaying. The as a surrogate for xMR.
stoichiometry of methane was used in these simulations. A typical Fig. 8 shows an example of such simulations for hydrogen, for
distribution of the normalized temperature rise above the initial the conditions of the experiment of Markides and Mastorakos
(inert) value Tin(x), defined as b ¼ [T  Tin(x)]/[Tb(xst)  Tin(xst)] with [103]. For these simulations, four different chemical mechanisms
Tb(xst) the burnt temperature at the stoichiometric mixture fraction, have been used [80,104–106]. Note that with detailed chemistry
is reproduced from Ref. [11] in Fig. 6. It is evident that ignition and variable cp, the initial distributions of species mass fractions
happens suddenly and that it takes a substantial time for the flame and the enthalpy will still be linear functions of x, but the
to acquire the expected shape for complete non-premixed temperature will not. Unity Le is assumed. It is clear that
combustion. In more detail, Fig. 7 shows the reaction rate, condi- a minimum exists; in this case and using, say, the mechanism of
tionally-averaged on the mixture fraction for various times during O’Conaire et al. [106], we would conclude that xMR ¼ 0.04 and
the induction period. It is evident that the reaction rate peaks at a x sref ¼ 1.05 ms. Fig. 8 also shows a sensitivity of the predictions to the
between 0.10 and 0.15, slightly shifting to higher values with time. detailed mechanism. Despite the expectation that detailed mech-
The value predicted by the analysis of Liñán and Crespo [10] for anisms of hydrogen are probably the best known, differences are
these conditions (of T1,0, T2,0 and Tact; these are the key parameters evident, although some convergence may be observed, with the
that affect the location of autoignition in mixture fraction space in more recent mechanisms producing increasingly similar results.
the analysis with one-step chemistry) is about 0.12. By reference to Second, we may use the asymptotic analysis results from Ref. [9]
Fig. 7, we may call most reactive mixture fraction, xMR, the value of x for constant-strain problems (also summarised in Ref. [49]) or from
where the reaction rate becomes a maximum. We see that xMR is not Refs. [10,90] for unsteady problems. Note that these are derived for
equal to the stoichiometric mixture fraction (0.055 for methane). one-step chemistry, but with a proper estimate of the activation
The simulations also showed an insensitivity of xMR to the turbulent energy they are applicable to real fuels as well (at least in
timescale, lengthscale, and initial mixing layer thickness. a temperature range where the autoignition of the fuel behaves in
The point raised above, i.e. that autoignition occurs away from an Arrhenius manner). Finally, we may perform laminar mixing
stoichiometry and always around the same mixture fraction that layer simulations and examine the location where the reaction rate
depends little on mixing rates and mixture fraction patterns, has peaks or autoignition occurs, e.g. from graphs as in Fig. 4. This
been confirmed with a wide range of additional DNS results with approach can fully incorporate differential diffusion effects and
simplified and complex chemistry, all of which reveal a similar some effects of strain [81,88].
structure of the autoignition sites in mixture fraction space when For the conditions of Ref. [11], xMR from the first method above
the fuel is colder than the air. Reported work includes simulations gave a value about 0.11, the asymptotic analysis gave 0.12, while the
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 67

Fig. 6. Scatter plot of normalized temperature above the initial inert value from 2-D DNS with one-step chemistry for various times until autoignition. The turbulence timescale sturb
in the original reference is such that autoignition occurs at about t/sturb ¼ 0.79. Reprinted from Ref. [11], with permission from Elsevier.

DNS showed a progressive shift from about 0.10 to about 0.15. researcher first calculates data as in Fig. 8, which not only can
Separate calculations of laminar unsteady mixing layers gave provide an estimate of where autoignition will occur in mixture
results that were very close to the DNS values from the turbulent fraction space, but can also identify the order of magnitude of the
calculations. We may hence conclude that the third method is the actual autoignition time, since sref will be determined simulta-
most accurate, as also suggested by Knikker et al. [88]. The neously. Note that performing a series of homogeneous auto-
stoichiometric mixture fraction is not directly related to xMR. ignition calculations with initial composition corresponding to
The determination of xMR is not very rigorous, as the peaks in a particular x, but without varying the temperature at the same
reaction rate or the minimum in sign are relatively broad (see, for time, may result in erroneous conclusions. If a transient laminar
example, the peak in temperature in Fig. 4 and the minimum in flame code is available, then full laminar flame simulations can be
Fig. 8). We also saw that xMR, if understood strictly as the x at which performed. In this case, caution is needed for the effects of scalar
the reaction rate peaks, may shift during the induction time and is dissipation and transport properties on the calculated autoignition
affected by scalar dissipation. With detailed chemistry, we must time [81,88].
also consider the fact that the various elementary reaction rates Mixing will usually delay autoignition from sref; there has been
peak at different mixture fractions. Despite these uncertainties, xMR no report so far that including mixing can decrease autoignition
is a useful concept, not least because it dissociates explicitly the time from sref. There is a chance that this may be so for hydrogen
autoignition mixture fraction from the stoichiometric mixture due to differential diffusion, which causes sign/sref to stay very close
fraction, which is normally where the reaction peaks during high- to unity in an unsteady laminar mixing layer for a range of initial
temperature combustion. It is suggested that when faced with an thicknesses of the fuel-air interface [88]. More laminar flow
unknown situation involving non-premixed autoignition, the investigations focusing on the ratio sign/sref are necessary to allow
68 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

8 945
7 Li et al.
Del Alamo et al.
6
O’Conaire et al.
Yetter et al. 940
5

ignition delay tign [ms]


4
935

T [K]
3

930
2

925

1 920
Fig. 7. Conditionally-averaged reaction rate against mixture fraction from one-step 0 0.02 0.04 0.06 0.08 0.1
DNS. The times at which the conditional averages were made, normalized by the mixture fraction Z [-]
resulting autoignition time for this simulation (with autoignition defined as the first
appearance of a burning spot), are 0.67 (open circles), 0.86 (filled circles) and 0.99 Fig. 8. Calculations of autoignition time in homogeneous hydrogen-air mixtures at
(triangles). Reprinted from Ref. [11], with permission from Elsevier. 1 bar with various detailed mechanisms: Li et al. [104], Yetter et al. [80], Del Alamo
et al. [105], and O’Conaire et al. [106]. The hydrogen mass fraction in the fuel stream is
Y1,0 ¼ 0.13 (the rest in N2), T2,0 ¼ 945 K, T1,0 ¼ 720 K. The initial species mass fractions
us to project with some confidence homogeneous calculations (that and temperature (dashed line; right axis) are functions of the mixture fraction cor-
responding to frozen mixing. sref is the minimum value (that depends on the scheme)
are done with modest computational cost) to inhomogeneous
and xMR the corresponding mixture fraction. Calculations performed by Frouzakis and
situations. Kerkemeier from ETH Zurich. Reproduced from [102].
If the fuel is distributed in the oxidizer in a manner that global
mixing may occur quickly and the extrema of x shrink so that the
nominal xMR (from one of the methods discussed above) ceases to temperature) kernels are located at mixture fractions leaner than
exist in the flow, the evolution of temperature in x space can be xst, although a quantification of xMR was not attempted. Later data
somewhat different than the above picture [95]. Nevertheless, the from a methane jet [111], with the co-flow temperature now at
fundamental finding that a competition between the high 1350 K, also showed that the highest temperature across mixture
temperature at low x and the high concentrations at higher x leads fraction space evolves from a very lean x to the stoichiometric with
to a preferred mixture fraction and that this value does not change downstream distance (Fig. 9), in qualitative agreement with the
much with flow configuration, remains valid. shapes from the examples of laminar flamelet simulations shown in
A situation more relevant to HCCI combustion, with autoignition Fig. 4 and the DNS (Fig. 6). In the methane case, it looks as if the
achieved by compression, has also been studied. DNS with one-step initial temperature increment occurs at very lean mixture fractions,
chemistry has shown that if the fuel and air have the same consistent with homogeneous autoignition calculations of sign in
temperature, autoignition happens around the stoichiometric that paper that show a minimum at an equivalence ratio of 0.05,
mixture fraction [107], although not many details were included for corresponding to xMR ¼ 0.01 for the jet dilutions used.4 The devel-
the temperature in x space. A 2-D simulation using a detailed opment between these lean mixture fractions and the stoichio-
scheme for hydrogen and including temperature non-uniformities metric, when examined in terms of conditional averages, seems
but uniform composition, showed the difference between local gradual (see also Fig. 5 of Ref. [82] for transient flamelet methane
autoignition and deflagration following a nearby autoignition site calculations at 40 bar demonstrating the same trend). These
[14]; this different behaviour has also been observed experimen- experiments have been the subject of various modelling efforts,
tally [108]. Earlier, Hasegawa et al. [109] reported similar localised which will be discussed later. Velocity data are also available [112].
autoignition in a fully-premixed gas with superimposed tempera- Fuels other than H2 and CH4 must also be examined in this
ture fluctuations with a 2-D simulation and one-step chemistry. configuration, perhaps following the hierarchy of the laminar flame
Strictly speaking, these should be considered as premixed autoignition studies, and the effect of pressure must be studied.
combustion problems. Development of appropriate diagnostics that can measure the
There is very limited experimental evidence on the autoignition mixture fraction for such fuels would also be very fruitful.
site structure in mixture fraction space. The ambient pressure
experiment by Cabra et al. at the University of Berkeley [110] 2.2.2. Low scalar dissipation
reported point Raman–Rayleigh and Laser-Induced Fluorescence The DNS shows that not all regions with x ¼ xMR autoignite
(LIF) measurements in a nitrogen-diluted hydrogen jet at about simultaneously [11]. Hence the autoignition process is ‘‘spotty’’ and
100 m/s issuing from a 4.57 mm pipe into a co-flow of hot products we can talk about local autoignition kernels. Such kernels have
from a lean hydrogen flame. The co-flow velocity was low (about been visualised in virtually all the DNS performed so far, with either
3.5 m/s) compared to the jet velocity, so the turbulence was created one-step or complex chemistry, in turbulent non-premixed flows.
by shear. The co-flow was at 1045 K and had an oxygen mole In accordance with our expectation from laminar flow auto-
fraction of 0.15. In the experiment, no extra ignition source was ignition, it turns out that the regions that autoignite first are those
available, but a lifted flame was evident, and hence this problem is that have low scalar dissipation. This was first revealed with one-
a proper turbulent non-premixed autoignition problem. Plots of
simultaneously-measured temperature and mixture fraction show
large scatter, with an overall gradual transition from the inert 4
Care is needed with the exact determination of autoignition times for such very
mixing line to the fully-burnt line over an axial distance of many lean mixtures that result in a very small temperature increment from the unburnt
jet diameters. The data suggest that autoignition (i.e. high to the burnt state.
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 69

Fig. 9. Scatter plots of temperature and OH mole fraction against mixture fraction from multi-scalar measurements from a methane jet in vitiated air at the indicated distance (z)
from the nozzle of diameter d. Reprinted from [111], with permission from Elsevier.

step DNS [11], but has been found with realistic chemistry as well. strained regions in the flow. The usual focus in turbulent mixing
Fig. 10, taken from Ref. [81], shows the localised nature of the studies is on the high N layers, which are aligned with highly
reaction rate along the xMR isoline from a 2-D hydrogen simulation strained regions [113]; such regions in the flow are not favourable
with detailed chemistry. High scalar dissipation can locally delay for autoignition.
the reactions due to heat and species loss and hence autoignition is The simulations with detailed chemistry of hydrogen and
retarded there. Low scalar dissipation, in contrast, allows the proper transport also show that the autoignition kernels are mostly
species and temperature to accumulate during the induction time, located in regions of low mixture fraction gradients [96,97] and in
which leads to autoignition. Regions with low N can be the cores of regions where the hydrogen is focused on the air side of the fuel-air
vortices, as suggested from 2-D [93,94] and 3-D [99] simulations by interface [96]; the effect of differential diffusion becomes impor-
correlating the fluctuations of N with an indicator based on the tant for hydrogen autoignition in turbulent flows, as in laminar
velocity gradient tensor that can distinguish between vortical and [81]. Echekki and Chen [97] also present a history of various radical-
70 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

of the conditional NjxMR seems to dominate: autoignition spots are


those that, in general, have experienced low NjxMR. (This symbol
means ‘‘N given that the mixture fraction is around xMR’’.) The
spatial statistics of the sites are determined by the spatial spreading
of the low NjxMR regions. The temporal and spatial statistics of the
conditional scalar dissipation fluctuations are therefore important.
Such information can be provided from experiment or DNS of inert
mixing, but there are problems, as discussed by Bilger [48], asso-
ciated with the resolution necessary to measure N and the rela-
tively low turbulent Reynolds number achieved in DNS. Further
work on turbulent mixing, focused on the low-N regions (and not
on the high-N regions as is usually done) and on conditional values,
is necessary.
There is yet no direct experimental verification for these DNS
findings. In a turbulent counterpart of the laminar counterflow
autoignition experiments, Kortschik et al. [59] used Planar Laser-
Induced Fluorescence (PLIF) of formaldehyde (CH2O) in a heptane
jet against heated air, with both streams being mildly turbulent
by the action of turbulent grids upstream of the stagnation region.
The creation of formaldehyde before autoignition has been dis-
cussed in the past for methane [116] and heavier hydrocarbons
[117]. Regions of high CH2O are located in the air side of the layer
and tend to be concave towards the air, i.e. at focusing points,
consistent with DNS.
Recently, an experimental set-up virtually identical to that of
Cabra et al. has been used at the University of Sydney [118] with
advanced diagnostics to probe the location and structure of auto-
ignition kernels. Simultaneous planar and quantitative measure-
ments of CH2O and OH (with PLIF) and temperature (with Rayleigh
scattering) have been performed in a natural gas jet into a slow co-
flow of vitiated air. The simultaneous CH2O and OH concentration
measurement can be used to provide a surrogate for heat release by
forming the product [OH]  [CH2O], following similar applications
to premixed flames (e.g. Ref. [119] and references therein). The
Fig. 10. Contours of heat release rate (left) with the xMR isoline superimposed from two
separate snapshots from 2-D detailed-chemistry DNS. The scalar dissipation along the flame has autoignited at a length from the nozzle that increases
xMR isoline is shown to the right. High heat release rates are found along the xMR isoline with decreasing co-flow temperature, as expected from an auto-
and at low values of the scalar dissipation. Reprinted from [81], with permission from ignition phenomenon. Fig. 11 shows typical simultaneous images at
Elsevier.
the base of this flame. The circled regions in the left column show
examples of fluid with low CH2O, low OH and a small temperature
filled kernels, some of which seem to be delayed significantly and increment; this may correspond to fluid approaching autoignition.
reach ignition late due to their exposure to local high scalar dissi- There are examples of autoignited fluid (visible OH, low CH2O,
pation during the induction time. Laminar unsteady starting jets visible reaction rate, high temperature) in the middle and right
[98] and laminar mixing layers interacting with vortices [92] also columns (circled). These regions have coincident high reaction
show a good correlation between the autoignition location with rates and low temperature gradients, something not expected from
positions along the xMR contours that have low scalar dissipation. a flame front, but expected from an igniting patch of fluid, if we
The good correlation between low N and high reaction rate is believe the DNS. The high-temperature gradients surround the high
further substantiated by calculating directly the conditional reaction rate spot. The boxes in Fig. 11 indicate measurements
correlation coefficient between these two quantities [11,93], which consistent with our expectations from a flame: high reaction rate
becomes very highly negative at the mixture fraction where the and high OH virtually coincide with steep temperature gradients,
reaction rate is maximised. Further data along these lines are which is the footprint of a reaction front, rather than an auto-
provided in Refs. [95,114]. The emergence of autoignition in regions igniting site.
of low scalar dissipation has also been mentioned in the A turbulent opposed-jet experiment with diluted hydrogen
compression-ignition simulations of Ref. [107]. versus heated air has been used at Princeton, with turbulence
It is evident that the fluctuations of the scalar dissipation are initially only in the fuel stream [120] and later in both streams
responsible for the spotiness of autoignition and this importance is [121]. As in the laminar counterflows, a critical air temperature Tcrit
probably greatest if N is generally high and close to Ncrit, since it is in above which autoignition does not happen can be measured. Tcrit
this parametric region that N will have its highest effect of auto- increases with bulk velocity, but it has a non-monotonic depen-
ignition time (consistent with laminar flame calculations with dence on turbulence intensity (controlled by the grid properties).
detailed chemistry, e.g. Fig. 6 of Ref. [51] for methane; Fig. 2 of For low velocity fluctuations (6% of the mean), Tcrit decreases
Ref. [68] for heptane). The autoignition kernel will happen in the compared to the laminar value, although it eventually increases as
minimum values of N in the turbulent flow. The effect of turbulent the velocity fluctuations become about 12% of the mean flow. A
straining, which increases the mean N, CND, but also its fluctuations decreasing Tcrit in the presence of weak turbulence has also been
N0 about CND, is not evident, since by increasing the fluctuations of N observed with heptane [59]. This trend is interesting and will be
we are also increasing the likelihood of obtaining low values (see understood best after examining in detail how autoignition times
Ref. [115] for a summary of results on the scalar dissipation fluc- depend on the turbulence, which is the topic of the next
tuations). In addition, in view of the importance of xMR, the history sub-section.
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 71

CH2O Only Small Kernel Medium Kernel


8x10-3

OH

1.5x10-3
CH2O

2200K
Temp

700K

1x10-5
RR

2000
K2.
(Grad(T))2

Pixel-2

Fig. 11. Simultaneously-acquired two-dimensional distributions of CH2O, OH, and T approximately at the base of a lifted methane jet flame in vitiated air. The reaction rate surrogate
[CH2O]  [OH] and the temperature gradient ((VT)2) were also calculated from the images. The flow was similar to the experiment from which Fig. 9 was taken. The jet centreline is
4 mm to the left of the image, which is 9 mm by 18 mm. Images supplied by Dr. R. L. Gordon and Prof. A. Masri of the University of Sydney [118].

2.3. Ignition times possible that the initial conditions of the simulation and the NjxMR
history control autoignition time, while for high sign/sturb, sign
In the previous sub-sections, it was shown that laminar and increases with sturb due to the slower mixing and hence the slower
turbulent simulations, with simple and complex chemistry, generation of well-mixed (low N) spots that promote autoignition
converge to the conclusion that autoignition kernels will lie along [99]. More work is needed in this area with detailed chemistry to
the xMR isosurface at locations with low scalar dissipation rate, i.e. form a unified picture of the various competing trends and it is
at the lowest NjxMR. The temporal evolution of the conditional important that the effects of initial conditions are always properly
NjxMR is therefore important. The initial thickness of the fuel-air presented and accounted for. The effect of non-unity Lewis number
interface, initial turbulent velocity and initial turbulent lengthscale has also been examined with 3-D simple chemistry DNS [95] and
seem to affect the first emergence of autoignition sites through significant alterations to the autoignition time and kernel structure
their action on the conditional scalar dissipation, as concluded by were observed.
examination of various 2-D simulations [11,93,96]. The DNS data In virtually all the DNS examined so far, the initial velocity field
show that the autoignition time sign (with sign understood as the was randomized according to a particular spectrum and the
first emergence of an autoignition kernel) is longer in all situations simulation then evolved with time. It is interesting to examine
where NjxMR is high. The ratio sign/sturb did not, directly, determine what happens if different realisations of the same flow are per-
how the autoignition time is affected by the turbulence and the formed. We are seeking information on the statistics of the timing
DNS data [11] show a relative independence of sign from sturb (sturb of first appearance of autoignition, for example on the mean CsignD
is the large eddy turbulent timescale). These findings may be and the r.m.s. sign,rms. This information may be useful for inter-
limited due to the fact that the DNS used was 2-D and hence the preting possible cycle-to-cycle variation in compression-ignition
mixing, and therefore the time of emergence of well-mixed, low-N engines and for assessing the likelihood of unwanted autoignition
spots could not be related properly to the turbulent timescale sturb. in a gas turbine premix duct.
However, they have not yet been discounted from 3-D DNS. The Unfortunately, very few data along these lines are available. The
effect of the initial lengthscale of the scalar distribution and the sign from six individual realisations of the DNS seem to follow a very
turbulence intensity have been closely examined in 3-D turbulence narrow distribution [81], with sign,rms/CsignD about 5%. In 3-D
[95,99] with initial x distributions from a turbulent spectrum. It is decaying turbulence and a straight fuel-air initial interface, the pdf
evident that in this case the mixing of x below (or above) xMR can of sign has been compiled over 25 simulations [122] (at a very low
have an important effect on autoignition time, while the initial turbulent Reynolds number, however) and the results show sign,rms/
scalar distribution and turbulent parameters have an important CsignD about 2%. Hence very little randomness is observed in sign
effect on the temporal evolution of CND and the conditionally- between different realisations of the DNS transient, although we
averaged scalar dissipation rate CNjxMRD. For low sign/sturb, it is cannot draw solid conclusions. More data focused on this aspect are
72 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

necessary. However, the small randomness of the time of first If sref becomes too high compared to the mixing timescales and
appearance of autoignition is consistent with the explanation that if the flow is such (e.g. as in an LPP gas turbine duct) that the
somewhere in a turbulent flow, the best possible history of NjxMR probability of encountering xMR is becoming small due to mixing
will emerge, and this spot will become the first autoignition kernel below this value, then sign may become very high. Such situations
[11]. have probably been observed in a confined flow turbulent auto-
The magnitude of sign in the turbulent flow relative to the ignition experiment at the University of Cambridge [103,127]. This
characteristic time sref is important to quantify. In all cases where experiment followed very closely the configuration shown in
this quantification has been made, it turns out that sign/sref > 1, Fig. 1a, with the turbulence created by a perforated plate in a duct,
which implies that we can understand sref as the minimum possible in the centre of which there was continuous injection of diluted
autoignition time in a non-premixed situation even in the presence gaseous fuel. The fuels tested were hydrogen, acetylene, ethylene
of turbulence. We do not expect sign/sref to be too high (ratios higher and pre-vapourised heptane. The air was electrically heated and
than 3 have not yet been reported in DNS) because this situation temperatures up to 1100 K have been reached at velocities up to
would correspond to a sustained high scalar dissipation rate about 30 m/s. The fuel velocity, U1,0 in the nomenclature of Fig. 1a,
everywhere. In the DNS of mixing layers, the autoignition time of could be higher than the air velocity U2,0 and so the flow would
turbulent flows was shorter than for laminar flows starting from correspond to a jet, or equal, so that the mixing corresponds to
the same initial condition [11,81]. In the configurations studied by a plume in approximately homogeneous isotropic decaying
DNS so far, mixing is manifested by a decrease in CND and CNjxMRD turbulence convecting at a uniform velocity U2,0. The overall
with time and hence eventually autoignition will occur, even if equivalence ratio was very lean (about 0.1) and the fuel nozzle was
NjxMR happened to be above the critical value initially. Ref. [93] smaller than the background turbulent lengthscale, suggesting that
contains data where the conditional mean CNjxMRD stayed above the the mixing process was dominated by the air flow turbulence.
critical value throughout the whole duration of the induction time Fig. 12 shows photographs from some of the resulting auto-
but still autoignition occurred. This was due to the fluctuations of ignition phenomena. These photographs are from the ‘‘random
the scalar dissipation, which resulted locally in spots with NjxMR spots regime’’ [103], where individual autoignition kernels are
lower than the critical value, and autoignition was possible there. evident. A random generation of autoignition kernels in various
The history of the fluctuations of the conditional scalar dissipation positions in the flow was observed. The photographs with short
was emphasized. Further DNS with the scalar dissipation sustained exposure times show individual flamelets (presumably grown
higher than the critical value for a long time compared to sref are around an autoignition spot) that do not give a connected flame.
necessary to elucidate these points further. The long-exposure photograph (Fig. 12c) shows a light-emitting
In the turbulent counterflow geometry [59,120,121], the flow region that is standing away from the nozzle at a well-defined
sustains the mean scalar dissipation constant in the radial direc- distance denoted as Xmin, implying the minimum distance observed
tion. Hence the emergence of critical conditions is a distinct from many instantaneous photographs. Under conditions of low air
possibility and indeed a Tcrit even in the presence of turbulence is temperature, there was no evidence of autoignition in the duct,
observed. It was mentioned in the previous sub-section that the which is probably related to the mixing to the overall very lean
data show that Tcrit in these experiments decreases at first over the mixture fraction determined by the relative air and fuel flow rates;
laminar value, but then increases again. This trend if fully consis- this condition probably corresponds to a very long sign/sturb. If the
tent with the DNS data. The earlier autoignition of the turbulent air temperature was very high, the autoignition spots appeared
mixing layer over the laminar mixing layer, as we said, has been
attributed to the fact that turbulent flows will create well-mixed
regions that have locally low N faster than laminar flows, creating
hence locally conditions that promote autoignition. A very intense
turbulence will tend to increase N so eventually the increasing
scalar dissipation and the associated retarding effect on auto-
ignition will dominate and Tcrit will increase again. The beneficial
effect of turbulence on ignition seems to be more pronounced for
low sign/sturb [96]. There is some evidence from experiment that the
development of the conditional scalar dissipation fluctuations
happens quite early (of the order of a few Kolmogorov timescales)
from the beginning of the mixing process [123], an observation
consistent with results on the generation of unconditional
scalar dissipation fluctuations from DNS [113]. Hence intense
turbulence may create low NjxMR regions quicker, which promotes
autoignition.
This trend has also been revealed in measurements of auto- XMIN
ignition times of a transient heptane spray injected in a weakly
turbulent flow as compared to the same injection in a laminar flow
[124]: in the turbulent case, the spray ignited earlier. Wright et al.
[84] discussed this experiment further through Conditional
Moment Closure (CMC) modelling that reproduces well the auto- 6 mm
ignition times for all conditions. Mizutani et al. [125] also observed
a reduction in sign of a column of droplets when the surrounding 50 mm
turbulence was intense. Kobori et al. [126] reported an autoignition
time that decreases with decreasing nozzle diameter in a diesel a b c
engine. Despite the fact that these experiments were done with
Fig. 12. Typical photographs of acetylene autoignition in the experimental arrange-
sprays, where extra phenomena are present, the results are ment of Refs. [103,127,135] (fuel continuously injected into hot co-flowing turbulent
consistent with the interpretations given in this sub-section on air). (a) Exposure time: 100 ms. (b) Exposure time: 250 ms. (c) Long-exposure
gaseous fuel autoignition. photograph of the flow of (a).
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 73

very close to the nozzle and eventually formed a continuous lifted A detailed analysis on the statistics of the autoignition timing
or attached flame, similar to the flames of jets in vitiated air and location from a diesel engine modified for optical access has
[110,111]; this situation corresponds to a low sign/sturb. The ‘‘random been presented by Baritaud et al. [131]. A large scatter of the
spots regime’’ shown in Fig. 12 was evident in the region sign/ autoignition location and timing was observed. The ratio sign,rms/
sturb z 1. Xmin decreased with increasing air temperature, as CsignD was about 20–30% for most of their data, except for some
expected, and increased non-linearly with velocity U1,0 and U2,0. For operating conditions of the engine at low temperatures or low
the experiments with U1,0 ¼ U2,0 and at the same air temperature, volumetric efficiencies that gave a ratio of about 50–60%. The
Xmin/U2,0 increased with U2,0. Due to the turbulent nature of the additional randomness introduced by the spray atomization and
flow and the distribution of residence times this implies, it is not evaporation would increase sign,rms/CsignD from a gaseous injection
clear if the ratio Xmin/U2,0 is an estimate of sign or of CsignD for value, so these diesel autoignition data are consistent with the
comparison with a transient injection problem; it provides, gaseous injection data of Ref. [129]. It would be very useful if
however, a measure of autoignition time that can be used to reveal experimentalists working with autoignition of transient jets
some trends. The lengthscale of the turbulence produced by the reported explicitly the statistics of sign so that the randomness of
grid and the ratio u/U2,0, with u the characteristic turbulent velocity the first autoignition time is better understood.
scale, did not depend on U2,0 and hence the delaying effect on Xmin/ We finally mention experimental results from impulsively-
U2,0 may be understood as due to the increased scalar dissipation started jets of high-pressure hydrogen from a small orifice, a situ-
associated with the high velocity. Modelling of this experiment ation that mimics an accidental (or intentional) release [40,132] or
with CMC [127] showed histories of CNjxMRD that were above the the bursting of a pipe. Here, the shock wave associated with the
critical value for a longer length along the pipe for the high velocity sudden release increases the temperature of the diffusion layer
case, which is consistent with a longer Lagrangian residence time formed at the interface between the hydrogen and the air, leading
until autoignition. Visualisation of individual spots from playing to autoignition. Very little analysis of this problem is available and
back long records of a high-speed film of light emitted from this the effects of flow (e.g. vortices, turbulence) are unexplored. Dryer
experiment helped the identification of the instantaneous auto- et al. [40] mention that localised autoignition may have happened
ignition length, Lign, which can be understood as the distance from in some of their experiments, but overall flame establishment was
the nozzle of the individual flamelet at the moment this flamelet is not achieved, which could be due to the high jet velocities. They
generated from autoignition. The resulting analysis gave Lign,rms/ also mention that the geometry downstream of the point of the
CLginD of about 15%. The available data to this day from this experi- leak is important, as it affects fuel-air mixing. It seems that the
ment include autoignition lengths as a function of various oper- phenomena of turbulent non-premixed autoignition described in
ating conditions, fuels, nozzle diameters and turbulence-promoting this review may have a relevance to this important problem also.
grids, and mixture fraction statistics and scalar dissipation
measured with PLIF of acetone in inert flow [123]. Further 2.4. Flame growth following autoignition
measurements are necessary in this flow to fully understand the
observed phenomena. Following localised autoignition, flamelets will grow around the
In addition to the continuous-flow gaseous autoignition exper- autoignition kernels, as evident in DNS that shows that reaction
iments of jets in vitiated co-flowing air [110,111,118] and the plumes zones propagate around the kernel [11,97]. The continuous-flow
in heated turbulent fast-flowing air [103,127], transient gaseous experimental arrangement of Markides and Mastorakos [103], from
jets have also been studied. Naber et al. [128] have examined the where the high-speed evolution shown in Fig. 13 is taken, also
autoignition of natural gas jets in a closed hot environment. showed this behaviour. However, there is very limited work on
Autoignition times were determined by the time when the pressure what really happens to these flamelets and how they evolve to
begun to rise from the start of the injection. The emphasis of this establish the whole flame.
work was on the kinetic effects of various species that may be The laminar flow simulations (e.g. Section 2.1) show that the
present in the natural gas and hence ignition locations were not rate of spreading of reaction fronts following autoignition increases
reported. Fast et al. [129,130] have used a similar arrangement for with the scalar dissipation. DNS simulations of turbulent mixing
gaseous dimethyl ether injection from a 2 mm nozzle. High-speed layers [11] have inferred the time needed for the whole layer to
imaging showed that the whole jet ignited at the same time. The reach burning conditions by monitoring separately the time
mean autoignition time was about 3 ms for the conditions reported, evolution of CTjxMRD and CTjxstD (with the angled brackets indicating
with an r.m.s. of about 0.5 ms, which give a sign,rms/CsignD of about conditional averaging over the whole DNS domain). The former
15%, i.e. larger than the preliminary data on the statistics of sign increases quicker than the latter, since, as we’ve seen, autoignition
obtained from DNS. This may be due to the higher turbulent Rey- occurs at xMR, which can be different than xst. The time shift
nolds numbers achieved in the experiments, due to background air between these two quantities, ssh, can provide some average
temperature fluctuations, or due to an inherent fluctuation in the information on the timescale necessary to ignite the stoichiometric
valve opening process. Laminar flamelet simulations suggested that contour. The quantity ½DðxMR  xst Þ2 =CNjxMR D1=2 was used to give an
the conditions of the experiment corresponded to a low strain rate estimate of the average distance between the xst and the xMR iso-
compared to the critical value and hence significant effect of surfaces, which then gave an estimate of the average velocity of the
straining on autoignition delay may not be present. It is interesting flamelets around the autoignition kernels. The result was that this
to analyse these novel experiments further. velocity was about 0.5SL, with SL the laminar burning velocity of

Fig. 13. Sequential images from a high-speed camera visualising the emergence of an autoignition kernel from the flow in Fig. 12. Arrow indicates flow direction. Reprinted from
[103], with permission from the Combustion Institute.
74 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

a stoichiometric mixture with reactants at Tin(xst). Alternatively, we located in regions of low scalar dissipation, further extending this
can write that ssh ¼ CDjxMR  xstj/[CNjxMRD1/2SL], with the constant C concept for situations with strong advection. The movement of the
about 2. Note that the number of individual spots distributed along spot in the predominant flow direction and its growth in all
a xMR isoline has little influence on this estimate. A large number of directions is evident (Fig. 14, centre and right columns), in good
spots (for example, if we had more high reaction rate islands along agreement with the experimental results. Eventually, flame fronts
the xMR isoline of Fig. 10) would imply a quicker rate of rise of the between individual autoignition sites collide. The process is
CTjxMRD and CTjxstD curves with time, but would not affect the delay statistically stationary. The fact that a point in space may be
between them, which is mostly determined by the physical experiencing at various times fully-burning or pre-ignition
distance between the mixture fraction contours. compositions is consistent with the scatter plots shown in Fig. 9.
This problem has also been addressed by Echekki and Chen The simulations capture qualitatively the spatial randomness of the
[97,133] with detailed-chemistry DNS. The authors suggest that the autoignition kernels and the various autoignition regimes seen in
time taken to establish the whole flame requires an additional the experiment. In particular, if the air temperature increases,
10–20% of the autoignition time. This transition is made through a connected lifted flame becomes evident in both the DNS and the
reaction fronts that first originate in lean regions (for their simu- experiment. Its base, however, is stabilised by autoignition: Fig. 15
lations xMR < xst), cross the stoichiometric line, and propagate into shows the build-up of HO2 ahead of the flame edge along the xMR
rich mixtures and along the stoichiometric isoline with a structure isoline and the increasing reaction rate there, characteristic of
similar to a triple flame. This structure has been reproduced in autoignition. Three-dimensional simulations are of course neces-
more recent DNS with imposed mean flow ([102]; see next sary to extend the scope of these conclusions.
sub-section). A planar hydrogen jet issuing with high velocity in hot slow air
We should also mention the interesting work of Seiser et al. has been simulated with 3-D DNS at Sandia National Laboratories
[134], who considered a laminar counterflow between hot air and [101]. In this configuration, the turbulence is produced by shear
a diluted hydrogen stream. Initiation of the flame was performed by between the jet and the co-flow and a continuous lifted flame is
a laser-induced photodissociation process that resulted in a flame established. In Fig. 16 a volume rendering of the OH and the HO2
kernel that grew and propagated along the stoichiometric mixture radicals is shown. HO2-containing regions extend way upstream
fraction isosurface. Unfortunately, the speed of this flame was not of OH-containing regions. The original uncertainty expressed by
measured. Further work on this problem can provide insight into Cabra et al. [110] whether this flame is stabilised by autoignition
the triple flame structures propagating into reactants that are close or flame propagation seems hence to have been answered by this
to autoignition. simulation that shows that the former phenomenon is mostly
A preliminary experimental validation of the DNS findings [11] responsible.
on the overall rate of growth has recently become available [135] In addition to the simultaneous presence of autoigniting spots
from the Cambridge turbulent autoignition experiment. In this and flame fronts shown in Fig. 11 from the vitiated air experiment,
work, the speed at which the flamelets grew around a kernel was which is qualitatively consistent with the DNS results with flow
measured by high-speed cinematography, an example of which is shown in Fig. 14, there are also some additional indications that the
given in Fig. 13. For a range of fuels (hydrogen, acetylene, pre- whole flame in these parabolic non-premixed situations is initiated
vapourised heptane) and flow velocities, it was found that the by autoignition. Simultaneous measurements of reactive scalars
average cross-stream speed of the flamelets was between 0.6 and and mixture fraction in a lifted hydrogen jet in cold air [136] show
1.2SL, with SL calculated from a laminar flame code with detailed that the transition of the conditionally-averaged temperature from
chemistry. This topic should be studied more extensively with the inert to fully-burnt curve (Figs. 9 and 10 in that paper) is
high-speed diagnostics and a wider range of mixing patterns. gradual. The peak conditional temperature and OH across x-space
always occur at xst. In contrast, the flames in hot vitiated air [111]
2.5. Flame stabilisation mechanism (their Figs. 9–11) show conditionally-averaged temperatures and
OH that peak at a mixture fraction that progressively shifts from
The presence of strong convection, in either a continuous-flow a lean value to xst with downstream distance. This is a major
device as in an LPP duct or in a transient jet as in a diesel engine, has qualitative difference between the two experiments, with the latter
raised questions on the mechanism of flame stabilisation and in behaviour fully consistent with DNS data of non-premixed
particular the role of flame propagation versus autoignition. autoignition.
Examination of the available DNS and experimental data suggest In the experiment of the hydrogen jet in vitiated air [110], with
clearly that it is the latter, as we will discuss in this sub-section. some experimental results also shown in the modelling paper of
Very few simulations of turbulent non-premixed autoignition Ref. [137], the lift-off height increased non-linearly with fuel jet
with mean flow have been performed. A two-dimensional super- velocity, contrary to the linear scaling expected from a lifted flame
sonic mixing layer has been considered [100], which shows that the in cold flow. In addition, it increased very steeply with a decrease in
onset of the flame is clearly due to autoignition. With smaller co-flow temperature. Decreasing T2,0 from 1080 to 1020 K increased
velocities, jets of hydrogen in hot air [101,102] have recently been the lift-off height from about 5 to 40 jet diameters. The change in
examined. These simulations show local fluctuations of the ignition laminar burning velocity in this temperature change is very small
point and demonstrate that autoignition is responsible for the (approximately 10% with a square dependence of SL on tempera-
stabilisation. In more detail, 2-D DNS of the Cambridge experiment ture), and hence a stabilisation mechanism due to normal edge
with hydrogen has been performed at ETH Zürich with a spectral flames [20] cannot account for this large change. In addition, the co-
element code with detailed chemistry and transport [102]. Extra flow was at a velocity close to the laminar burning velocity of
care was needed to inject in the DNS domain the right turbulence. a stoichiometric mixture, which is just about the limiting condition
Fig. 14 shows snapshots of the heat release rate from this simula- for blow-off of a lifted flame in cold co-flowing air [138]. In the
tion, indicating an individual autoignition spot emerging upstream Cambridge experiment, the co-flow was between 2 and 5 times the
of a flame-like region. The spot is visualised with a small ‘‘hole’’ laminar burning velocity of a stoichiometric mixture at these
created in a region of high HO2 concentration and lies close to the conditions, depending on the fuel used. Edge flames do not prop-
xMR isoline, which was pre-computed from the data of Fig. 8, while agate at such high speeds (we will discuss this in more detail in
the flame includes reaction zones along the xst isolines and at rich Section 3) and hence in that experiment too the combustion was
and lean mixtures. Analysis of the data shows that the kernels are clearly stabilised by autoignition.
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 75

Fig. 14. 2-D DNS of the experimental arrangement of Ref. [103] with detailed chemistry and transport for hydrogen. The flow comes from below and is turbulent and the air
temperature corresponds to the ‘‘random spots’’ regime. The hydrogen jet is in the middle. Top: logarithm of heat release rate; bottom: HO2, with OH isolines superimposed in
purple. Times: 0, 0.06 and 0.13 ms from left to right (the origin of time is not important). The xMR ¼ 0.04 and xst ¼ 0.17 isolines are marked with the dashed and the solid line,
respectively. Reproduced from Kerkemeier et al. [102].

This discussion has a bearing on diesel engines, and in particular The presence of formaldehyde upstream of a hydrocarbon
on the observation that the flame is often at a well-defined flame initiation point suggests an autoignition-driven stabilisation
stand-off distance from the nozzle [139] if injection of fuel is still mechanism. This has also been observed in a jet of methane in
occurring after autoignition. Extrapolation of the above findings to a highly diluted preheated air flow [140], a characteristic situation
that situation implies a small effect of transport processes (i.e. for MILD combustion [31].
flame propagation) and a large contribution of autoignition of the To complete this sub-section, we summarise conceptually in
fuel-air mixing layer as it flows towards the flame base. The tran- Fig. 17 the various timescales involved in the process of flame
sition from autoignition to the conventional diffusion flame diesel establishment following autoignition in a turbulent non-premixed
combustion mode passes through the so-called premixed phase that flow, as revealed from the experimental and DNS results reviewed
involves flamelets originating from autoignition kernels that move here. Time could be understood as a Lagrangian time or as distance
along and across the mixture fraction contours. An example of the in statistically-steady problems or as the time from the beginning of
propagation along the contours is shown in Fig. 15 from DNS (see injection in transient problems. Averaging could be understood as
also Figs. 2 and 3 of Ref. [133]). We saw that propagation occurs in volume or ensemble averaging, depending on the situation. The
fluid that is already ready to autoignite, which results in a very chemistry appears mostly in sref and, together with the operating
quick flame expansion process. An example of propagation across conditions, in xMR, while the scalar dissipation rate and its statistics
mixture fraction contours is given by the movement of the transi- affect all the other timescales. Further work is needed for the
tion between the inert and the burning lines in T–x scatter plots statistics of sign, as we have seen. The rates of rise of CTjxMRD and
from DNS (e.g. Fig. 6) or experiment (e.g. Fig. 9) towards rich x. This CTjxstD are affected by the number of autoignition sites and the flame
process is completed faster with increasing scalar dissipation. expansion along the mixture fraction contours and there is no clear
These points are discussed further in the context of CMC simula- picture yet on how turbulence affects these. Finally, very prelimi-
tions by Wright et al. [84], although for situations that do not result nary information is available for the time delay ssh, which is related
in a lifted flame. The relative speeds of deflagration and autoignition to flame spreading across x-contours. Comparing modelling results
fronts in a uniform mixture with temperature non-uniformities, with experiment should be done based on the same quantity and
a problem of indirect relevance for the first flame expansion mode Fig. 17 can help identify what exactly is being compared. Note that
discussed here, has been discussed in Refs. [14–16]. the evolution of unconditional averages in autoignition is
76 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

Fig. 16. Volume rendering of OH (left) and HO2 (right) from the DNS of a planar
hydrogen jet in hot air [101]. The flow comes from below, the hydrogen jet is in the
middle of each image. Courtesy of Dr. J. H. Chen of Sandia National Laboratories, image
produced by the assistance of Prof. Kwan-Liu Ma and Hongfeng Yu of the SciDAC
Ultrascale Visualization Institute at UC Davis.

2.6. Modelling approaches

In this sub-section, we discuss various modelling efforts for


capturing turbulent autoignition. Most of these have been applied
to diesel-like problems, with few so far applied to continuous flows.
Most modelling efforts are based on Reynolds-Averaged-Navier-
Stokes (RANS) modelling, although recently Large Eddy Simulations
(LES) have also been performed. We expect that the latter will
become more widespread with time and as the sub-models become
more reliable. Despite the fact that many of the modelling efforts
concern sprays, the two-phase element of the model is usually not
considered in detail. Some comments along these lines are also
given here.
The average behaviour of the flow should be understood as
a time or ensemble average for continuous-flow problems that are
statistically steady, such as the spatially-developing mixing layer or
the jet (Fig. 1). For transient problems such as the diesel spray
injection, the average should be understood as an ensemble over
many realisations. In DNS of temporally-evolving flows with
homogeneity in at least one spatial direction, averages have been
compiled over the homogeneous direction(s) at a particular time
instant. Due to the usual thinking in terms of autoignition time, in

<T|ξst>
τsh

<T|ξMR>
X

Fig. 15. As in Fig. 14, but with the air temperature corresponding to the lifted flame
regime. Left column: t ¼ 0, right column: t ¼ 0.03 ms later (the origin of time is not <τign> X
important). Top: HO2 mass fraction; middle: OH mass fraction; bottom: the logarithm T2,0
of the heat release rate. Note the autoignition development ahead of the flame edge. Tin,st
Reproduced from Kerkemeier et al. [102].
Tin,MR
τref τign
Time
complicated by the fact that the pdf of mixture fraction may also be
changing with distance or time. Since, for example for the Fig. 17. Conceptual presentation of the various timescales involved in the autoignition
R1 of a turbulent non-premixed flow for the case T2,0 > T1,0 and xMR > xst used as example
temperature, CTD ¼ 0 CTjhDPðhÞdh, trying to understand auto- (xMR < xst could also happen). sign is meant to indicate one autoignition time realisation
ignition only through the evolution of CTD may hide important from an ensemble that gives the average CsignD. For an estimate of ssh based on DNS
trends. data, see Section 2.4.
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 77

autoignition problems with continuous flow the above distinction premixed problem if both the temperature and the species have
must be borne in mind. finite fluctuations. The best way to understand the effects of mixing
frequency on autoignition time in these homogeneous situations is
2.6.1. Mean reaction rate probably through the notion of a xMR compared to the long-time
Early attempts to model autoignition in diesel engines with CFD (well-mixed) mixture fraction of the particular flow studied.
completely ignored species segregation and solved equations for The hydrogen jet flame in vitiated air has been simulated by the
the average mass fractions of species and temperature with transported scalar PDF method [110,137,150] and the joint
a chemical source term evaluated at these average values. Macro- frequency-composition PDF method [151]. The methane flames have
mixing was included by the usual gradient hypothesis for the been simulated so far only by joint-scalar PDF [111,152,153], to the
turbulent scalar flux and even detailed chemistry has been used. author’s knowledge. The common conclusions from these simula-
However, micromixing was not included. This option is still avail- tions are that, in this flow, the chemical mechanism plays a major role
able in some commercial CFD codes and these approaches have re- in predicting the right lift-off height, the mixing model changes the
surfaced lately due to the interest in modelling HCCI-type of diesel shape of the T–x scatter plots but does not alter much the prediction
engines, where the injection happens early and hence significant of the lift-off height, and that the flow is stabilised by autoignition,
mixing is achieved before autoignition. Clearly, such approaches are with HO2 generated ahead of the hydrogen flame and both HO2 and
fundamentally wrong. As soon as gradients of species mass fraction CH2O ahead of the methane flame. Large sensitivities to the co-flow
exist, fluctuations will be generated in a turbulent flow and hence composition and in particular to H2 and OH are also found [152]. The
there is an inconsistency in solving averaged equations with spatial sensitivity of hydrocarbon autoignition to NO that has been revealed
transport using a mean reaction rate that neglects the species in laminar flow autoignition [73] should also be considered.
fluctuations. The accuracy of results achieved with this method is, The transported PDF method has also been used to predict the
as expected, not good, as discussed by extensive computations and critical temperature in the turbulent opposed-jet hydrogen auto-
comparisons with other models in Ref. [141]. ignition experiments of Ref. [121] and it seems that the inclusion of
Modern variants of the Eddy Dissipation Concept (EDC) [142] differential diffusion is important [154]. The authors attribute some
have been used to give the mean reaction rate in compression- discrepancies with experiment to the molecular mixing model in the
ignition engines [143]. Cabra et al. [110] have also used this model PDF equation. Indeed, it is precisely for this critical condition, where
for their hydrogen jet in vitiated air experiment, with further the mixing rates are so high compared to the chemistry that auto-
results [144] exploring the sensitivity of the predictions to the ignition may be precluded, that the mixing model will have the
boundary conditions. The small-scale inhomogeneities and highest importance. In view of the finding that autoignition sites are
detailed chemistry are included in this model. However, there is no those that occur in regions with low scalar dissipation discussed in
detailed comparison on whether this model can predict the Section 2.2.2, it may be suggested that PDF methods with some
delaying effect of high scalar dissipation on autoignition time. In measure of the mixing rate fluctuations should be more appropriate
principle, it may be able to predict the flame expansion following for non-premixed autoignition. Indeed, Cao et al. [151] conclude that
autoignition by a propagation mechanism, as it is considered the joint frequency-composition PDF method gives better agree-
applicable for both premixed and non-premixed combustion. ment with experiment than the joint-scalar PDF method. An
Further exploration of this model and its application to experi- Extended Interaction by Exchange with the Mean (EIEM) model has
ments that show a distinct effect of turbulence on autoignition is also been proposed [155] as a method to include the fluctuations of
necessary before solid conclusions can be drawn. mixing rate in the joint-scalar PDF method and promising results
were shown for autoignition of a spray in a closed vessel.
2.6.2. Transported PDF In practical calculations for HCCI engines emphasizing the effect
The transported PDF method is probably the first turbulent of initial conditions and subsequent mixing on the onset of auto-
closure approach that has been applied for autoignition problems, ignition [156–158], the joint-scalar PDF method has shown good
and autoignition is probably the first turbulent reacting flow results compared to global experimental observables such as the
phenomenon that this method has been applied to. Dopazo and pressure trace. In these examples, we do not expect significant
O’Brien [12] considered the autoignition of binary reactants in the effects of small-scale strain on autoignition. Such a situation would
limit of high activation energy and assumed a constant correlation probably make the engine operation prone to large fluctuations in
coefficient between the species concentrations. They focused on ignition timing which was not reported in the experiments that
the temporal evolution of the temperature pdf and showed that, as these modelling efforts were trying to capture.
expected, the initial width of the temperature pdf affects the time In the context of LES, both the Berkeley and the Cambridge
of autoignition, since high-temperature fluid particles will auto- autoignition experiments have been modelled [159–161] with
ignite first. Mixing during the induction time, however, can a sub-grid PDF solved by the Stochastic Fields method. This is a new
homogenise these fluctuations and autoignition (defined as the approach that is derived from the one-point PDF equation and
time when the average temperature rises) may be retarded. Further holds promise in reducing the computational cost. The results for
work in such homogeneous (in the mean) problems includes both experiments suggest autoignition as the flame initiation
Monte-Carlo solutions of the joint-scalar PDF from the initial mechanism. In the case of the duct flow hydrogen experiment
condition to complete combustion. Relatively slow one-step [103], the LES [160,161] captures qualitatively the autoignition spot
chemistry compared to the mixing timescale was first considered evolution shown in Fig. 13 and quantitatively the autoignition
[145] and a wider range was then examined [146,147]. The results length Xmin (Fig. 12). For the hydrogen jet in vitiated air, very good
showed a substantial change of the pdf shape with Damköhler agreement with experimental data in all respects is reported [159].
number. In [147], the pdf method was also applied to a starting jet It would be interesting to examine more flow conditions, for
of fuel in hot air, predicting autoignition at the sides of the jet. The example to try to capture autoignition length as a function of flow
effects of sprays on the PDF equation have also been considered velocity, in order to see whether the model can indeed capture the
[148]. The case of spatially-homogeneous pdf with finite width was small-scale straining effects on autoignition.
also examined by Correa and Dean [149] with a Monte-Carlo
method. They considered uniform temperature and non-uniform 2.6.3. Laminar flamelet
species and, in this case, ignition was retarded as mixing was faster. The laminar flamelet model has been used extensively for
These findings cannot be generalised to a generic autoignition non- autoignition in diesel engines since the mid-90’s. In this model,
78 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

transport equations for the mean mixture fraction and for the autoigniting reactors to LES of the methane flame of Cabra et al.
variance of the mixture fraction fluctuations are solved in the RANS [111], including new models for the progress variable in the context
code and a presumed pdf of the mixture fraction, P(h), usually taken of LES, provides very good results [174]. The relative importance
as a b-function, is used to give the mean species mass fractions and between autoignition and local flame propagation at the sub-grid
temperature. The input to this integration, apart from P(h), is the scale is also inferred from these simulations. The model, in its
distribution of the reacting scalars against mixture fraction, which present form, probably cannot account for the possibility of strong
is calculated by simultaneous solutions of the flamelet equations local effects of scalar dissipation on autoignition, but it is a prom-
(Eqs. (2) and (3)) that provides data such as those in Fig. 4. Detailed ising framework where such effects may be added in the future. It is
chemistry of diesel surrogates such as n-heptane is often used. For interesting that the hydrogen flame in vitiated air [110] has been
more details, see [47]. Various choices on how N(x) is modelled captured successfully with LES using tabulated species mass frac-
and on how the flow-field information from the CFD code can be tions from pre-calculated perfectly-stirred reactors [175]. We
used to provide the scalar dissipation needed in the flamelet wouldn’t expect such tabulation to give good results in situations
equation have been explored. In principle, the flamelet equations with autoignition delayed due to strain, which possibly suggests
can be solved in every point of the CFD grid and hence the evolution that in this experiment such effects were not strong.
of each one will depend on the local value and history of N. In
practice, a smaller number of flamelets is used, to give the so-called 2.6.4. Conditional Moment Closure
‘‘Representative Interactive Flamelet’’ (RIF) model. RIF has also The Conditional Moment Closure is a mixture-fraction based
been used for natural gas jet injection and subsequent autoignition method that solves modelled transport equations for conditionally-
[82]. A development, called the ‘‘Eulerian Particle Flamelet Model’’, averaged reacting scalars, the conditioning done on the mixture
where the flamelets are presumed to exist with a probability I in fraction. For the species mass fraction of species a, the CMC equation
the flow with I coming from an advection-diffusion modelled for the conditionally-averaged mass fraction Qa ¼ CYajh ¼ xD is [176]:
equation, has been proposed to take mean flow effects into
   
account [162]. ~ hÞ v00 Y 00 h
vQa 7, rPð a v 2 Qa
With these flamelet models, the effect of scalar dissipation on þ hvjhi,7Qa ¼  þ hNjhi 2
autoignition is directly taken into account, as is the second phase of vt ~ hÞ
rPð vh
flame propagation (across the x isolines) following autoignition. þ hwa jhi (4)
Propagation along mixture fraction isolines and the effects of the
fluctuations of the scalar dissipation, however, are not included. To with h the sample space variable of x. Modelling is necessary for the
improve on the latter point, flamelet models with fluctuations of N conditional scalar dissipation rate, the conditional velocity CvjhD, the
included have been developed [163,164] based on a presumed log- conditional turbulent flux Cv00 Y00 ajhD, the Favre mixture fraction pdf
normal pdf of N. The comparison with DNS results is reasonable, ~ hÞ, and the conditionally-averaged reaction rate CwajhD. A similar

but application to a multi-dimensional problem has not been equation can be derived for the enthalpy or the temperature. The
attempted yet. To improve on the former, which may be important explicit presence of the scalar dissipation makes the CMC model
in diesel engines with multiple injections where the flame gener- a very suitable candidate for capturing the effects of small-scale
ated by autoignition from a first injection event will spread into mixing on autoignition. The fact that pre-ignition reactions may
a fuel spray injected later, a flamelet model with two mixture cover a relatively large range of mixture fractions, and the fact that
fractions (one for each injection) was developed [165]. A justifica- in a turbulent flow the scalar dissipation may have strong varia-
tion for neglecting the cross-scalar dissipation between the two tions in mixture fraction space, makes the explicit inclusion of CNjhD
mixture fractions was given. It would be interesting to see more in CMC a very desirable characteristic. The CMC framework allows
extensive use of these new developments. also the derivation of transport equations for the conditional
Pre-computed laminar flamelet methods [166–169] have also fluctuations. These equations include explicitly the conditional
been used for transient problems, with a progress variable or correlation between the fluctuations of a scalar and the scalar
time being a second parameter for the tabulation in addition to dissipation, and hence are very appropriate for capturing
the mixture fraction. Some success is evident and, if care is taken the important effect of the fluctuations of N on autoignition. Finally,
in the modelling of the source term in the progress variable the presence of the convection and the turbulent diffusion terms
equation, the stand-off height of the flame in a long-duration in the governing equation suggests that the physics of flame
injection diesel spray can be calculated accurately [169]. Such expansion along mixture fraction isolines can be included. There-
models should be tested against more experimental data, and in fore, CMC is eminently suitable for autoignition and hence it has
particular from simpler parabolic problems such as the autoigniting been used for various problems so far.
jets described here previously. Kim et al. [177] used first-order CMC, meaning that
The Flame Surface Density (FSD) model is related to the flamelet CwajhD ¼ wa(Q1, Q1,., QT), for the natural gas transient injection
and is hence described here. Various variants [170–172] have been experiments of [128]. The CMC equation was used in an integrated
developed, making use of chemistry tabulation. Contrary to the way form, which makes the CMC equation (Eq. (4)) similar to the
the flamelet model is implemented, the FSD model solves equations transient flamelet equations (Eq. (3)), and in two spatial dimen-
for the flame surface per unit volume that include various mech- sions. Reasonable agreement with experiment was found and some
anisms of flame production and destruction validated by DNS. Such differences between the elliptic and the integrated models were
models have shown good agreement of autoignition time in observed, indicating an effect of physical transport to the auto-
a transient jet [172]. Further application of these models would be ignition time prediction. Similar efforts for transient spray auto-
interesting. ignition in constant-pressure vessels include Refs. [84,178], while
Tabulation is particularly attractive for LES due to the low the first-order CMC model has also been used for 3-D simulations of
computational cost burden it imposes on the already expensive diesel engines [179]. Overall, good agreement with data was shown
flow solution, while allowing for some phenomena associated with despite the neglect of the conditional fluctuations. This has been
finite rate chemistry to be included. Further validation of the explained by the relatively low CNjhD compared to the critical value
tabulation methodology for laminar non-premixed autoignition [84], making hence the conditional fluctuations having a minor
problems has been reported [173]. Application of such tabulated effect. A variant of the CMC method, which uses pre-calculated
methods based on premixed flamelets and homogeneous transient laminar flamelet results to provide a closure for CwajhD
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 79

called Conditional Source Estimation [180], has also been used for transport of the conditional averages [185]. Multi-dimensional CMC
transient methane jet autoignition. has also produced very interesting insights [186] into the stabili-
Second-order CMC, where the conditional reaction rate includes sation mechanism of the hydrogen flame in vitiated air [110].
the conditional fluctuations [176], has been applied to autoignition Doubly-conditioned CMC models, where the reacting scalars are
and validated against various DNS databases [114,181]. The Cam- conditionally-averaged on the mixture fraction and a progress
bridge autoignition experiment with pre-vapourised heptane has variable, are also available [95], but these have not been used in
also been modelled with the first-order [127] and the second-order realistic problems yet. This approach may provide a good way for
CMC [182] giving very good results for autoignition lengths. In this predicting the flame expansion process for fronts of type 3 in Fig. 2.
experiment, the second-order correction to CwajhD makes a small The CMC model has also been used in LES of autoignition [187].
contribution because, apart from a small region close to the nozzle, Fig. 18 shows the conditionally-averaged temperature from various
the scalar dissipation rate decays well below the critical value and axial positions from the Cabra et al. methane experiment [111] and
hence the conditional temperature fluctuations are small. These from simulations with LES, with CMC used as the sub-grid turbu-
become large, however, as the autoignition point is approached. The lent combustion model. Excellent agreement is found. The LES/CMC
second-order CMC model is expected to be very useful in situations combination for autoignition problems seems very encouraging in
with high scalar dissipations. Ref. [181] shows that increasing the that it may capture both the locality of autoignition and the effects
value of the ratio CN00 2jhD1/2/CNjhD in the model in situations where of molecular mixing on it. In addition, LES can capture the move-
CNjhD/N(h)crit > 1 can cause the system to switch from a frozen state ment of reaction zones due to large-scale motions, e.g. by large
to autoignition, in agreement with DNS. Hence, the second-order eddies in a jet. This may have an influence on the stabilisation
CMC model can capture many of the physical aspects of turbulent mechanism of the jet flames in vitiated air studied experimentally.
autoignition described in this paper. It would be interesting to LES-CMC may be used as a thorough tool by which important
apply second-order CMC to the turbulent counterflow autoignition questions can be explored.
experiments [59,121] that show a critical condition, although
calculating accurately the scalar dissipation in this flow may require 2.6.5. Other
the solution of a modelled transport equation [183]. Echekki [188] has also used the Linear-Eddy Model with detailed
The model used for the turbulent flux term Cv00 Y00 ajhD is based on hydrogen chemistry for a globally-homogeneous fuel-air flow with
a gradient hypothesis. However, counter-gradient diffusion has small-scale inhomogeneities. The model produces useful informa-
been observed in DNS of propagating turbulent edge flames by tion on the chemical structure of autoignition, as it reveals the
Richardson et al. for low values of u/SL [184]. Although this topic has effects of molecular mixing on various elementary reactions. The
not been examined for autoignition problems yet with DNS or model has great potential to be included in RANS simulations or LES
experiment, it is plausible that the high temperatures in auto- of realistic flames.
ignition problems imply a relatively low u/SL. We estimate u/SL in Danby and Echekki [189] have taken a very different approach to
the range between 0.2 and 4 for the conditions of the Cambridge the usual strategy of providing a model in the context of RANS or
and Berkeley experiments at the base of the flame, which implies LES. These authors have analysed DNS data through the Proper
that elliptic CMC modelling for these problems may necessitate Orthogonal Decomposition method. Despite the fact that the
attention to this term. The model for this term is also important for usefulness of this method is not fully established yet, it may offer
capturing flame expansion in multiple-injection diesel engines. a new way of looking at chemistry-autoignition interactions.
Despite these comments, the conventional model has produced From the point of view of chemical operation safety, Gourara
very good results for a diesel engine with a pilot injection before the et al. [190] suggested the calculation of residence times with LES by
main injection, a situation that critically depends on spatial performing particle tracking and comparison of these times with

2380

x/D=30 x/D=40
1860

1340

820

300

x/D=50 x/D=70
1860

1340

820

300
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4 0.5
Mixture Fraction Mixture Fraction

Fig. 18. Conditionally-averaged temperatures from the experimental data of Ref. [111], shown in Fig. 9, and from LES-CMC simulations. Solid line: Multi-dimensional CMC; dashed
line: cross-stream integrated CMC. Note the shifting location of the peak conditional temperature from lean value to the stoichiometric mixture fraction. Reproduced from Ref. [187],
with permission from the combustion Institute.
80 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

a characteristic autoignition timescale such as sref. This comparison process (probably a small effect in the DNS due to the small droplet
can be used to assess autoignition hazards in chemical mixers. This size) and the random position of the droplets in the flow. Although
approach is not really suited for predicting autoignition for not studied in detail, the structure of temperature vs. x before
combustion applications, but it introduces some of the concepts we autoignition seems similar to that of gaseous fuel autoignition
discussed here to a different and important area. (Fig. 11 of Ref. [210]), with a xMR about 0.025 for an air temperature
T2,0 ¼ 1500 K. Similar data have also been revealed with 3-D DNS, but
2.7. Spray autoignition with single-step chemistry [211]. The evaporation reduces the
temperature of the xMR regions, which has a delaying effect on
Compared to our knowledge on how turbulence affects the autoignition time. Further simulations with droplets in turbulent
emergence of autoignition sites in gaseous fuel problems, there is flows are necessary with 3-D DNS codes and detailed chemistry.
virtually no information on this topic available for sprays. In In the context of modelling, the turbulent combustion models
a turbulent spray, the fuel evaporation introduces an additional described in Section 2.6 have all been used for spray autoignition,
timescale to the problem and the droplet size and spacing intro- e.g. for diesel engines, with good results obtained in general, as long
duce additional lengthscales. These parameters influence auto- as the fuel generation process is modelled correctly and at least the
ignition in a complicated manner, also manifested in the simplest mean mixture fraction is captured with sufficient accuracy [212].
problem of a single droplet suddenly immersed in a hot environ- However, little attention has been paid to how some features of the
ment. In this situation, which has been examined extensively models may need to be explicitly altered due to the presence of
analytically [191], numerically [192–198] and experimentally evaporation. The mixture fraction source due to the fuel generation
[199,200], the autoignition time increases with increasing droplet is straightforward to include in the equation for CxD and this is
diameter and decreases with increasing fuel volatility. For a review always done in diesel engine simulations. The transport equation
of the literature prior to 1998 and a presentation of the key trends for the variance of the mixture fraction is also affected by evapo-
with temperature and pressure for various fuels, consult Ref. [8]. ration and some models are available [213,214], but their validity is
For hydrocarbon fuels at air temperatures and pressures of rele- not fully established yet. The mixture fraction pdf in the presence of
vance to applications, the autoignition location is at a radius large evaporation has also been considered [148,213,215], but again has
compared to the droplet diameter [196] and corresponds to lean not received enough attention. There has been very little work on
mixtures [197,198]. There is some evidence that the Negative the influence of evaporation on the scalar dissipation [216,217] and
Temperature Coefficient (NTC) region of hydrocarbon autoignition, more needs to be done in this area. Due to the difficulties in
where the autoignition time of homogeneous mixtures is reduced measuring accurately the mixture fraction in evaporating sprays,
as the temperature increases, may be absent in droplets in infinite there is little guidance from experiment on the magnitude of the
hot surroundings due to the vapour flow [194]; however, a two- neglected phenomena. Our turbulent reacting flow theories need to
step temperature rise has been experimentally observed [201]. It be expanded to include consistently droplet evaporation effects
would be interesting to explore single-droplet autoignition simu- and our experimental techniques for measuring mixture fraction in
lations in the context of xMR and sref that characterise non-premixed two-phase flows must be improved.5
gaseous autoignition. Such comparison may be possible with the
data of Refs. [197,198]. 3. Spark ignition of turbulent non-premixed flames
Autoignition of sprays in a hot air turbulent flow has been studied
as early as 1949 by Mullins [202] in the context of gas turbine reheat In this section, we will discuss fundamental findings for the
systems. In this experiment, a kerosene spray injected in a uniform forced ignition of turbulent non-premixed flows. Despite its prac-
turbulent flow of hot air has shown CH2O chemiluminescence before tical significance, this is a new topic of research in turbulent
the autoignition point and cool flames at low air temperatures, non-premixed combustion and hence has not been studied as
consistent with modern diesel engine studies [117]. Mullins [203] extensively as autoignition. However, new technologies such as the
also reviewed early work on spray autoignition and reported that the gasoline direct-injection spark-ignition engine and lean gas
autoignition delay time in a flow reactor increased with an increase turbines have some difficulties with ignition and hence this topic
of the Sauter mean diameter of the spray. has recently become a focal point of research. We will start by
A numerical model for the autoignition of clusters of droplets in recalling some relevant results from studies of ignition of fully-
convective flows [204] has suggested that autoignition may be premixed systems, which have, in contrast, been examined
happening around individual droplets or around the cluster as thoroughly and we will continue with forced ignition of laminar
a whole depending on the inter-droplet spacing relative to the non-premixed flames. The probabilistic nature of spark ignition of
droplet diameter, with some interesting experimental results from turbulent non-premixed flames and the subsequent flame expan-
a droplet cluster recently showing similar behaviour [205]. Aggar- sion are then discussed. We will close with a discussion of
wal [8] discussed the various droplet cluster autoignition regimes modelling efforts and some comments on spray ignition.
in detail and compiled data from various experiments. Even
without flow, the presence of fuel vapour around a droplet alters 3.1. Insights from premixed systems
the time and location of autoignition [206]. Simulations of laminar
flow around droplet arrays at low Reynolds numbers show that the 3.1.1. Ignition sources and the minimum ignition energy
evaporation rate and drag of each droplet is different from single The most widespread method of initiating combustion of
droplet behaviour [207]. There seem to be no explicit data on how a flammable gas is by an electric spark, while other methods
these results can be safely extrapolated to turbulent flows. include laser ignition, flame jets and plasmas. Laser ignition has
2-D DNS of spray autoignition in turbulent flow, with the droplets been reviewed by Ronney [218], which could also be consulted for
taken as point sources and hence with the inter-droplet region not its succinct review of spark ignition, with some more details on the
resolved, have only recently been performed [208–210] using flow structure around the laser spark discussed in Refs. [219,220].
complex chemistry and relatively high air temperatures so that the Ignition by a flame jet [221] or puff [222] has been suggested to
NTC regime was not evident. The results show a strong influence of
droplet diameter on autoignition time and on the ‘‘spotiness’’ of
autoignition kernels, which is now influenced additionally by the 5
Developments on CMC including droplet evaporation are under way at the
randomness introduced by the turbulence on the evaporation University of Sydney. Refs. [211,217] also include some relevant data.
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 81

improve ignitability of the mixture due to the increased initial by the gas due to heat losses that depend on geometric parameters
surface area of the kernel. More recently, a detailed study [223] of a and on the proximity of the flame kernel to the electrodes. The
transient hot gas jet into a quiescent hydrogen-air mixture showed experiments of Lewis and von Elbe have been correlated approxi-
that the first site of ignition appears at the tip of the hot gas get. mately with MIE w d2q , with dq being the quenching distance for the
Plasma jets have been proposed as effective igniters for gas turbines mixture [6]. Analytically, various estimates have been proposed
[17,224] and their action is not only thermal, but also chemical, based on the balance of heat generation inside the kernel with the
since they release significant amounts of radicals. A situation called heat loss to the unreacted gas ahead of the kernel. Glassman [4]
‘‘radical-induced ignition’’ has been observed in numerical simu- gives that MIE ¼ (4/3prc3)rcp(Tb  T0), where rc is the critical radius
lations [225] of autoignition of uniform mixtures starting from of the spark and Tb the adiabatic flame temperature. A simplified
cold temperatures with large amounts of radicals present initially. (but insightful) early analysis [203] gives that rc ¼ 3l/(rbcpSL). A
If their concentration is such that they do not decay due to more complicated analysis [232] suggests that the critical flame
recombination reactions faster than they can trigger the generation kernel will be the one that contains in the burnt gases energy equal
of more radicals, the system will eventually autoignite. to the energy stored in the preheat zone of the expanding kernel;
Ignition in the mixing layer between hot products of combus- this results in a critical Karlovitz number condition and reproduces
tion and a cold premixed fuel-air flammable mixture, which may approximately the d2q dependence of the Lewis and von Elbe
mimic locally forced ignition by a large pilot or the stabilising effect experiments. Other experiments, however, are closer to a d3q
of a recirculation zone behind a flameholder, has been analysed in dependence [7]. All experiments show that rc and dq are several
1954 by Marble and Adamson [226]. They found that ignition times greater than the laminar flame thickness.
occurs well into the hot gas stream and a front slowly propagates For methane, the MIE across the flammable range is a minimum
into the flammable mixture stream. In this flow, before any at a slightly lean equivalence ratio, while for heavier hydrocarbons
significant heat release, the temperature profile follows the well- the MIE is a minimum at rich mixtures (e.g. see data in Glassman
known error function shape as expected from unsteady heat [4]). Hence the MIE is not necessarily a minimum where the
conduction. The ignition point was associated with the emergence laminar burning velocity is largest. This is attributed to preferential
of a point with zero cross-stream temperature gradient, indicating diffusion [232] across the highly curved flame front when the
a local temperature peak [226]. More modern thinking of this kernel is in its early stages.
situation has revealed the so-called ‘‘Homogeneous Charge Diffu-
sion Ignition’’ regime [72], where the high temperature of the hot 3.1.2. Spark ignition of turbulent premixed flows
gas stream may not really allow us to talk of a sudden transition Increasing the flow velocity past the electrodes increases the
from an unburnt to a burnt state, but rather of a gradual progres- MIE because the flow stretches the spark downstream, which
sion. The reaction rate in a counterflow between hot products and increases the spark path, and hence causes the energy input to be
ultra-lean reactants was finite across all the ‘‘mixture fraction’’ distributed over a much larger volume that reduces the tempera-
range (with the mixture fraction here defined as unity in the hot ture reached [4]. If the flow is turbulent, there is an additional heat
products and zero in the air-fuel mixture) and depended addi- loss mechanism from the kernel due to the higher strain and hence
tionally on the strain rate [72]. This has also been discussed through a larger MIE is needed, with ample evidence for this provided by
experimental observations [71]. If, therefore, the region of hot gases early [233] and later [234,235] experiments. More recently, Huang
available to initiate combustion is infinite and their temperature is et al. [236] reported systematic experiments of MIE of uniform
high, combustion will always begin. methane-air mixtures at equivalence ratio 0.6 from a fan-stirred
Sparks are usually not large compared to the flow and hence cruciform burner for a range of u/SL and found that MIE w (u/SL)0.7
additional considerations are necessary. An electric spark involves in the range 0 < u/SL < 23, with MIE increasing from 2.14 mJ to
many processes that are quite fast relative to the slow flame 10.15 mJ. They also reported a steeper rise of MIE with u/SL for u/
propagation process it induces. The initial breakdown phase opens SL > 23 and attributed this to the fact that the flow reached the
a conducting channel between the two electrodes, in which distributed reaction zones regime. Their data is reproduced in
subsequent current flow deposits energy to the gas. This energy Fig. 19. A model due to Ballal and Lefebvre exists, based on an
quickly increases the temperature of the gas and shock waves extension of the MIE estimate presented in Section 3.1.1, but with
emanate from the sparked region at a timescale of the order of ms the critical kernel radius explicitly determined as the quenching
from the beginning of the spark; the high temperature leads to distance. The latter has been appropriately altered to include
ignition of the gas between the electrodes [6]. Various investigators turbulent flow [237]. The complete model of Ballal and Lefebvre
have solved numerically the system of governing equations for this that includes expressions for spray forced ignition is summarised in
initial phase [227–231] and included the energy deposition by the Ref. [238]. It is not immediately evident if this model reproduces
plasma. The results show the initial phases of shock wave formation newer experimental data such as those of Huang et al. [236].
and propagation, a torroidal flame kernel, the transition to an A separate interesting feature of the experiments with spark
approximately spherical flame, the sensitivity to electrode shape ignition in turbulent premixed gases is that the phenomenon is
and gap, and reveal the effects of spark energy and duration. The probabilistic, in the sense that ignition was not achieved every time
general picture that emerges is that after all the fast processes have a spark was deposited in the turbulent mixture. So, in the experi-
subsided (after a time of the order of 100 ms, depending on the ments of Huang et al. [236] with u/SL ¼ 10 for example, an ignition
spark), in successful ignitions a flame kernel a few mm in diameter probability of 50% was obtained when the spark had an energy of
has been created. This type of analysis is very relevant to under- about 4 mJ, but 100% was reached (i.e. all spark events were
standing the various timescales of the spark-ignition process. successful) for a spark energy of about 10 mJ (Fig. 19). This differ-
For most practical purposes, the overall energy deposited by the ence is larger than the stochasticity expected from the reproduc-
spark is the most important factor. Measurements reveal that there ibility of the ignition system used in these experiments, since
exists a minimum ignition energy (MIE) for initiating a self-sus- similar measurements in laminar flows showed a narrower range of
tained front. If the energy deposited is too low, the flame kernel MIE between the 50% and the 100% ignition probability. This
decays rapidly because heat and radicals diffuse out of the kernel implies an additional randomness in the success of ignition,
faster than their rate of production inside the kernel. The MIE presumably due to the effect of the turbulent fluctuations of the
depends on parameters such as electrode diameter and gap and the strain rate at the moment of the spark at the spark location. Note
spark duration because not all of the electrical energy is absorbed that this can be safely inferred from the fact that these authors did
82 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

ξrich

ξst

ξlean

xlean xst xrich x

T
L

Tsp

T0

Fig. 19. Minimum ignition energy measurements for a uniform methane-air mixture xsp xlean xst xrich x
in isotropic turbulence. The single triangle denotes the MIE for stagnant mixture.
Reprinted from Ref. [236], with permission from the Combustion Institute.
δsp

not observe flame quenching after some initial flame growth and Fig. 20. Schematic of the parameters appearing in the forced ignition of a non-pre-
hence ignition failure can be associated with the kernel behaviour mixed flame. Upper: mixture fraction distribution. Lower: The thin lines denote
possible temperature profiles during the fast combustion initiation at the spark posi-
immediately after the spark energy deposition ceases. The fact that tion, if we think of the spark as a localised heat source. The thick line denotes
the ignition is a probabilistic phenomenon has also been reported a possible temperature profile once the initial fast transient has decayed.
by Akindele et al. [25], who also measured in detail the flame kernel
radii as a function of time separately for failed and successful flame. Intuition suggests that in contrast to the forced ignition of
ignitions. The conclusion was that failure occurs quite early, when a homogeneous mixture, here the spark’s location relative to the
the flame kernel radius was 2–3 mm, interpreted as the radius at positions of the stoichiometric mixture fraction xst and of the
which the strain rate of the expanding spherical kernel falls below nominal lean and rich flammability limits xlean and xrich, must play
a critical value [234]. The initial rate of growth (up to a time a role in the success or not of the ignition process. Hence, if the
approximately equal to the spark duration) did not depend strongly distance of the spark from the stoichiometric position L ¼ xst  xsp is
on the turbulence because the kernel was too small to experience very large, the heat produced by the spark will decay before any
the whole range of eddies, with the kernel just being convected by appreciable reaction picks up. The width of the spark dsp will also
the flow. A recent interesting exploration of kernel failure by Klein play a role, since a wide spark may overlap with some of the
et al. [239] was based on DNS data of spherical kernels analysed in flammable region, causing ignition at lean or rich mixture fractions
terms of the displacement speed, which was found to decrease and that can then ignite the whole flame. Even if the spark is inside the
even be negative (i.e. the reaction zone receding relative to the nominally flammable region, i.e. xlean < xsp < xrich, if the tempera-
diffusion of fuel in the preheat zone of the flame) when the ture resulting from the initial fast sparking transient is not high,
turbulence was intense. A ‘‘spark-assisted propagation’’ regime was then again the hot region will decay without causing ignition.
also observed [234] where the flame kernel growth was assisted by ‘‘Ignition’’ here is defined as the emergence, at long times, of a non-
the large energy of the spark. The reader is cautioned to the fact premixed flame. The presence of strain will alter the rates of these
that the flame thickness, appearing in both MIE expressions and in processes and can hence change the limits of ignitability. We
interpretations of data from initial flame growth, may not always be conclude that the spark position, width, energy and the strain rate
defined in the same way, which alters some numerical results but can affect the ignition of a non-premixed flame.
not the trends discussed here. A systematic study of these effects has not been performed yet.
The possibility of kernel quenching in turbulent flows is Rashkovsky [243] may have been the first to analyse the problem of
important for understanding misfire in spark-ignition engines. flame ignition in a strained counterflow mixing layer between cold
Failed events can be attributed to lean mixtures at the spark loca- fuel and cold air by a source of energy at a given location across the
tion, possibly caused by mixing of the fresh charge with trapped layer. All the energy Esp was delivered at a point; molecular diffu-
gases from the previous cycle, which hence necessitate very high sion then caused the region of high temperature to spread.
energy to ignite [240]. The cycle-to-cycle variability in stratified- One-step chemistry was used and a constant-strain field was
charge engines may also be related to the mixture composition at assumed. Detailed examination of the results of numerical calcu-
the spark location at the spark instant [241]. However, misfires also lations revealed important trends. First, that there is a critical
seem to occur even if the mixture is fully homogeneous. This is Damköhler number for successful flame establishment. Second,
consistent with the probabilistic nature of ignition discussed that this critical Damköhler number increases sharply with
previously and has additionally been attributed to the possibility of decreasing Esp and, third, that it increases sharply with increasing
a bulk kernel movement by the turbulence towards the spark distance of the spark position from the location of the stoichio-
electrodes or the combustion chamber wall [242]. metric mixture fraction (in Rashkovsky’s simulations, xst ¼ 0.5 and
therefore the resulting flame resided at the stagnation plane).
3.2. Spark ignition of laminar non-premixed flames The above trends have been confirmed with laminar counter-
flow flame simulations of methane with the GRI3.0 mechanism
Fig. 20 shows schematically some of the parameters involved in [244]. In these simulations, the initial condition for the scalars was
the forced ignition of a laminar non-premixed one-dimensional the inert distribution corresponding to the set strain rate, while the
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 83

temperature was set at 293 K everywhere except for a slab of gas There are very few experimental studies focused on the forced
that was set at Tsp, as shown in Fig. 20. The use of detailed chemistry ignition processes of laminar non-premixed combustion. Laser-
and transport allows a better representation of the reaction initi- induced ignition kernels in laminar hydrogen [245,246] and
ation at both lean and rich mixture fractions, while the use of methane [247] jets have been visualised to determine ignition
flame-derived mechanisms for this problem is reasonably reliable success and the positions along the jet that result eventually in
as long as Tsp is not too high. Fig. 21 shows the evolution of the successful jet flame establishment. Similarly to premixed flames,
temperature with time for three sparks, differing only in their the kernel success depended on whether the initial kernel around
position across the mixing layer which was at a strain rate less than the laser spark could undergo transition to a propagating flame,
one half of the extinction strain rate. It is evident that not all sparks although details on the critical spark energy were not given. These
are successful. Failure may occur at strain rates Scr,sp lower than the experiments also showed that the overall ignition of the jet flame is
extinction strain rate Sext of the non-premixed flame we are trying a composite problem, comprising both the initial generation of
to ignite. In Fig. 2 of Ref. [244], quantitative data on the ratio Scr,sp/ a non-premixed flame locally and the flame propagation along the
Sext are given that show that this ratio depends on Tsp and on spark jet to initiate the whole flame. The former phase is what has been
position. Despite the fact that spark C in Fig. 21 was placed at the considered in the one-dimensional simulations [243,244] pre-
stoichiometric position while spark B was close to the lean limit, C sented above. The latter phase involves triple or edge flame prop-
failed while B was successful. This was attributed to the scalar agation [247]. The propagation speed of laminar edge flames has
dissipation rate distribution in the opposed-jet flow, which received considerable attention and some results will be discussed
increases as x increases at the low x around stoichiometry consid- later to compare with new findings on turbulent edge flames.
ered here. Flame establishment involves reaction zones travelling The numerical predictions described above on the critical strain
across the layer to consume the premixed fluid, similarly to what rate for non-premixed flame ignition have not been observed yet in
has been observed following autoignition of non-premixed layers. experiment. Measurements of MIE as a function of strain rate and
Here, however, the speed of these fronts is smaller due to the spark position in a laminar strained flame would be very interesting.
completely frozen nature of the fluid ahead of them.
These simulations confirm the complicated picture of forced
ignition of non-premixed flames. In view of the lack of experi- 3.3. Ignition probability: flammability factor, Pker and Pign
mental data on this topic and the challenging physics involved, it is
considered that this area is virtually unexplored. Many more As we saw in Section 3.1, forced ignition of a turbulent homo-
simulations are necessary for various pressures and temperatures, geneous mixture involves additional complexities to the ignition of
since the operating conditions alter the flammability limits and will a stagnant mixture of the same equivalence ratio: (i) extra spark
hence alter Scr,sp/Sext. In addition, detailed studies of the initial energy is needed to ensure ignition in order to overcome the
spark processes of the type performed for homogeneous mixtures, detrimental effects of the higher strain rates associated with the
such as those in Refs. [227,229,231], must be performed for situa- turbulent flows, and (ii) a certain degree of stochasticity is expected
tions with inhomogeneous fuel distribution in the electrode gap. due to the strain rate fluctuations at the spark position and location.
In addition, as we saw in Section 3.2, forced ignition of laminar non-
premixed flames can fail if the strain rate is too high compared to
a critical value, which depends significantly on the spark energy
and location. It is hence evident that sparking a turbulent non-
3000 2.0x10-7s
Stoichiometry

premixed flow will include these phenomena and additionally be


Spark A 2.0x10-6s
2000 2.0x10-5s influenced by the mixture fraction fluctuations at the spark loca-
2.0x10-4s tion. In the context of Fig. 20, we may envisage that turbulence will
1000 2.0x10-3s make the distance L into a random quantity and the fluctuating
Time→∞ turbulent strain rate will cause fluctuations in the distance
0 xrich  xlean. Hence, the success of a sparking event to result in
Spark B a flame kernel involves a great deal of stochasticity and some
Temperature (K)

3000 results on this are given next. We distinguish between the proba-
bility of igniting a kernel, Pker, from the probability of establishing
2000 the whole flame, Pign.

1000
3.3.1. Jets
The probability of initiating a flame kernel from a spark in
0
a turbulent axisymmetric jet has been measured for a range of fuels
Spark C
3000
in a pioneering series of experiments by British Gas [23,248,249].
The creation of a kernel was judged visually. In these experiments,
2000 the electrode gap was 3 mm and the spark had energy 100 mJ,
which was higher than the MIE of homogeneous mixtures over the
1000 whole flammable range. The spark duration was 20 ms. Natural gas,
propane and simulated town gas (50% H2, 30% natural gas, 20% N2
0 by volume) were examined. Mixture fraction measurements were
0 0.1 0.2 0.3 0.4 0.5
performed with Raman scattering [23] and the data were analysed
Mixture Fraction (Z)
to give the pdf of the mixture fraction, P(h), at every point. From
Fig. 21. Transient temperature profiles in a laminar non-premixed counterflow layer this, the flammability factor was defined from
between methane and air with a spark represented by a slab of gas set at Tsp ¼ 4000 K Z xrich
at t ¼ 0. For all, the strain rate was 0.4Sext, with Sext the extinction strain rate. The F ¼ PðhÞdðhÞ (5)
centre of spark A was located in a region leaner than xlean, B was at xlean and C was at xlean
xst. B was successful, while A and C were not. Simulations performed with a laminar
flame code and the GRI3.0 detailed mechanism. Reprinted from Ref. [244], with The experiments showed: (i) that the region where Pker > 0 extends
permission from Taylor and Francis. further than the region bracketed by the isosurfaces CxD ¼ xlean and
84 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

CxD ¼ xrich, implying that the mixture fraction fluctuations play had a negative correlation with ignition success (see Fig. 14 of
a role; and (ii) that Pker z F along the centreline of the jet. Ref. [251]), with failed events associated with high positive
With ‘‘ignition’’ defined as the initial kernel generation plus (downstream) velocities and successful events with lower veloci-
successful flame propagation upstream to ignite the whole jet ties. The limiting downstream distance, beyond which upstream
flame, the probability Pign (called ‘‘conditional light-back proba- flame propagation in the jet does not occur, has been determined
bility’’ by Smith et al. [249]) can be different to Pker. The results also by using a pilot flame as the ignition source [252]. These
showed that Pign sharply reduces to zero downstream of a certain authors called this distance ‘‘Upper Propagation Limit’’. This limit is
distance. For example, for natural gas issuing at 20 m/s, Pign equivalent to the downstream location where Pign falls to zero in
decreased from unity to zero between 80 and 100 jet diameters, Fig. 23. For a range of jet and co-flow velocities, the data are
while Pker decreased gradually to zero at about 180 jet diameters. correlated with the estimated mean flow speed at the stoichio-
The results also showed a reduction in Pign with jet velocity. In good metric surface. The existence of a downstream location beyond
agreement with our expectations from the self-similar behaviour of which ignition does not result in a flame but simply in kernels that
an axisymmetric jet, F did not change with velocity. It is not fully are blown-off has also been observed in tests concerning the
clear from the data whether Pker depended on the jet velocity or ignitability of hydrogen leaks [253].
not. The overall findings from these studies have been confirmed It is not clear yet whether the relationship Pker ¼ F holds also for
from experiments with large-scale natural gas releases with drive off-axis locations. The concept that Pker ¼ F is reasonable if the
pressures up to 100 bar [250]. spark energy is higher than the MIE over all flammable range (it is
Similar data for a methane jet, including radial distributions of in all experiments described above) and if the spark can be
Pign and the effects of spark parameters on Pign have been reported considered as a point in space and in time. Ahmed and Mastorakos
in Ref. [251]. Fig. 22 shows snapshots from a high-speed movie [251] reported an increasing Pign with increasing spark duration for
showing the flame propagation process from this experiment. the same energy. Given that increasing the spark duration increases
Failure of the overall process was associated either with failure to the probability of sampling flammable mixture at the spark’s
initiate a flame kernel (no visible flame after the spark ended) or location, this trend can be explained. It is not clear yet if local strain
with a kernel that was convected downstream. Once a flame started can cause Pker < F, although some evidence that points to this
propagating upstream, it was likely to propagate all the way possibility exists from ignition in other geometries (discussed
upstream to the stabilisation point. These observations are later). Further work on directly correlating Pker and F in the jet,
consistent with the earlier studies [249]. Contours of Pign are employing detailed diagnostics at the spark location and instant,
reproduced in Fig. 23. It is evident that: (i) Pign decreases sharply would be fruitful. Simultaneous imaging of the reaction zone and
downstream of the estimated position where the mean stoichio- scalar dissipation at the spark location would be most useful.
metric mixture fraction of the inert jet closes on the centreline;
(ii) Pign < 1 even along the mean stoichiometric isoline; and (iii) Pign 3.3.2. Counterflow
is reduced with increasing velocity. Spark ignition in the counterflow geometry has been examined
It is understandable that Pign is reduced when the jet velocity is by Ahmed et al. [254]. The geometry was as shown in Fig. 1c and the
high, since the flame kernel will have to propagate against a higher spark was placed at various positions across the mixing layer and
velocity to ignite the whole jet. The local velocity measured by along the radial direction. The counterflow allows good control of
Laser Doppler Anemometry at the spark at the moment of the spark the bulk strain rate and the experiments can therefore shed some

Fig. 22. Flame evolution following spark ignition at 40 jet diameters on the centreline in a jet issuing methane partially-premixed with air (30% by vol.) from a 5 mm nozzle at
25.5 m/s. Reprinted from Ref. [251], with permission from Elsevier.
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 85

Fig. 23. Ignition probabilities (defined as successful kernel generation plus upstream flame propagation) as a function of spark position. The jet velocity was 12.5 m/s for the left
contour and 25.5 m/s for the right. The estimated CxD isolines corresponding to the lean and rich flammability limits and the stoichiometric mixture fraction are superimposed.
Reprinted from Ref. [251], with permission from Elsevier.

light on the strain effects and on the differences between F, Pker and flame grew and filled a significant part of the combustor including
Pign. It was found that Pign decreased with increasing bulk velocity U some of the recirculation zone, but overall flame stability was not
even when the spark was on the centreline. Since when sparking at achieved. These three failure modes correspond directly to the
the centreline the radial convection assists flame spreading three phases of non-premixed flame ignition mentioned in the
(contrary to the jet where the flame must propagate against the Introduction. High-speed cinematography showed that successful
flow), a reduction in Pign with an increase in U implies a reduction of overall ignition was associated with flamelets that moved upstream
Pker with increasing U, which is consistent with localised quench- while in the central region of the flow, presumably assisted by the
ing. The maximum bulk velocity for which a flame could be ignited recirculating flow there.
was about 90% of the bulk velocity of extinction of the stable flame. The data showed that: (i) Pker < F in general, so that Pker was as
In the language of the laminar flame simulations of Section 3.2, low as 20% inside the recirculation zone where F was almost unity;
Scr,sp/Sext was around 0.9. It is interesting to note that this ratio (ii) Pign < Pker or Pign z Pker depending on the flow location; (iii) the
agrees almost quantitatively with the extinction strain rate of flow rates of fuel and air alter all these distributions considerably;
a travelling edge flame relative to the extinction strain rate of an and (iv) approximately, the most ignitable region is the shear layer
established non-premixed flame (see Section 3.4.3), implying that between the annular flow and the recirculation zone at about the
loss of ignitability in this case may also be associated with impos- maximum width of the recirculation zone. Imaging of flame kernels
sibility of edge flame propagation along the layer following with OH-PLIF inside the recirculation zone suggested that failure of
a successful kernel. As for the jet, it is not yet clear if Pker ¼ F in all the kernel at its early stages is similar to our expectations from
positions and at what bulk or turbulent strain rate we would get kernel failure in premixed combustion, since the recirculation zone
Pker to fall to zero. is well-mixed. In particular, the measurements of mixture fraction
In these experiments, F was determined by employing Eq. (5) on and velocity suggested that the ratio u/SL inside the recirculation
measurements of the mixture fraction pdf performed with PLIF of zone was close to the limiting value for quenching of a premixed
a tracer (acetone) laden with the fuel. It was found that Pign could be flame at the local equivalence ratio. There was no detailed imaging
greater than F along the centreline. Hence, even if the spark was of the failed kernel in regions with steep mixture fraction gradients.
located in a region where F ¼ 0 (such as deep into the fuel or air These experiments demonstrated that F, Pign and Pker have
streams), ignition could sometimes be achieved. The reasons for complicated distributions in a realistic combustor and that
this are the finite size of the kernel (2–3 mm in diameter) following knowledge of F alone is not sufficient to determine the locations
the fast expansion during sparking, and a long-range effect, asso- across the flow that are appropriate for forced ignition.
ciated with the convection of heat or radicals created at the spark to Minimum ignition energies for laser ignition in a swirling
the flammable region, causing ignition there [254]. The possibility heptane flame, with significant degree of pre-vapourisation, have
that transport may have a beneficial effect on forced ignition has been measured by El-Rabii et al. [255], who also reported Pign at
also been seen in laminar flame simulations [244], where sparking various positions as a function of the laser energy. The presence of
far from the lean or rich limits could still ignite the flame (provided liquid droplets in this burner complicates the interpretation of the
that the strain rate was not above Scr,sp). The opposed-jet flow results in the context of clarifying the differences between F, Pker
revealed both local and non-local effects on the success of kernel and Pign, but the data suggest strong spatial variations in Pign. Laser
generation and subsequent flame propagation. It is expected that ignition in a hydrogen burner [256], with simultaneous measure-
these complications are evident in other flows as well. ment of the mixture fraction using a spectroscopic examination of
the spark kernel (a technique called Laser-Induced Plasma Spec-
3.3.3. Recirculating flames troscopy [257]), showed again a strong spatial variation of Pign and
A detailed comparison between F, Pker and Pign has been little correlation between Pign and the local x at the moment and
attempted in a bluff-body stabilised non-premixed methane flame location of the spark. No comparison between Pign and F was made.
[22], with a flow pattern similar to that shown in Fig. 3. As in the jet, The LIPS technique may be very suitable for a simultaneous
the two most common modes of failure were (i) failure to initiate measurement of F and Pker or Pign, although the ‘‘local’’ x given by
a kernel (no light visible after the spark finished) and (ii) creation of the measurement is the average over the volume of the spark,
a kernel which is then convected away. A third mode of failure has which is not negligible. More measurements on this topic and for
also been observed in this, more complicated, geometry: (iii) the this geometry are necessary.
86 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

3.4. Flame growth following forced ignition we can call it non-premixed edge flame. This distinction is consistent
with that suggested by Bilger et al. [19]. In view of Eq. (5) that
In order to achieve good ignitability of the flame, we need to defines the flammability factor F, we may call non-premixed edge
generate the kernel (Pker > 0) and to allow this kernel to expand in flames those in regions with F < 1, while premixed stratified flames
order to ignite the whole combustor. The various flame propagation those with F ¼ 1. A flame in a mixture fraction pdf around stoichi-
phases were shown qualitatively in Fig. 2 and discussed in the ometry with F ¼ 1 (denoted as S1 in Fig. 24) could also be thought of
Introduction. Some information on this topic is discussed in this as an edge flame, but the relatively small mixture fraction fluctu-
sub-section, with some further comments on the overall flame ations in this case would probably suggest that such flames are best
ignition for the practically-important flames stabilised by recircu- understood from premixed flame concepts. A flame such as S3,
lation zones. Failure associated with these propagation phases will which includes mixtures with compositions leaner than xlean and
result in Pign < Pker. hence has F < 1, is mostly of premixed character. Combustion in
mixtures with x < xlean in this flame may be facilitated by diffusion
3.4.1. Flame classification from the vigorously-burning mixtures with x > xlean, which implies
Müller et al. [258] presented some experimental results for the that we should treat xlean and xrich only as nominal values. The
average flame position as a function of time in a methane jet ignited distinctions discussed here are mostly of a qualitative nature.
at a downstream location. The flame stabilised at the lift-off height Turbulent flame propagation in both types of stratified mixtures is
and these measurements were used to validate a model for flame an important constituent of the problem of forced ignition of non-
propagation in stratified mixtures, with the model including premixed combustion. At present, it is not clear if the non-pre-
concepts from both premixed and non-premixed combustion mixed flame left behind at xst after the travelling front has passed
theory. Later, similar measurements for more spark positions were through a region with a pdf S1 or E makes any subsequent contri-
performed [251] in order to increase the amount of data available bution to the overall ignition process. In the bluff-body stabilised
for model validation. OH-PLIF and line-of-sight images of the flame non-premixed flame [22], in some ignition events the flame was
after sparking at the jet centreline (Fig. 22 is an example of the seen to de-stabilise and blow-off even after the flame had grown to
latter) showed that the flame first expanded quite quickly across the ignite the whole flow; the reason is not clear but it may be related
jet. The slower process of upstream propagation begun after about to a failure of the non-premixed flame. Simultaneous mixture
10 ms from the spark, at a time when the flame kernel was almost fraction and temperature measurements at various stages during
20–30 mm in diameter. Sparking at an axial distance of 40 nozzle the flame expansion and the stabilisation phases are necessary to
diameters D (D ¼ 5 mm) in a jet with 12.5 m/s of a CH4-air mixture elucidate the exact nature of these flames.
(air 30% by vol.) resulted in a transient of about 400 ms before the
flame stabilised at a lift-off height h/D ¼ 5, while a flame in a jet 3.4.2. Turbulent stratified-charge premixed flames
with 25.5 m/s took 700 ms to reach the lift-off height of h/D ¼ 15. Chapter 4 of Peters’ book [47] contains an extended review of
Measurements of the mixture fraction [248] showed that the the literature until 2000 on flame propagation in stratified
flammability factor from Eq. (5) was very close to unity around the mixtures. A model is also provided based on the G-equation and
stoichiometric isoline on the centreline of an axisymmetric jet, but a turbulent flame speed that relies on SL(x) and the pdf of x, that
decreased upstream. Hence, following a successful spark, the flame apparently gives good results for the stabilisation height of a jet
travels in areas with very different F. In a bluff-body flame, sparking flame [259]. More recently, experimental data on flame propaga-
inside the recirculation zone or close to its boundaries [22] also tion in mixtures with inhomogeneities have become available from
involved a first propagation phase in mixture that was mostly simplified geometries [260–262] and realistic direct-injection
flammable and had small mixture fraction fluctuations (i.e. F z 1) spark-ignition engines [263]. These results, together with DNS
and a later expansion in the whole flow in regions with F < 1. findings [264], suggest that the inhomogeneity introduces an extra
In an effort to classify the nature of flames involved in the wrinkling mechanism that can make the turbulent flame speed
transition from a kernel to the fully-established non-premixed higher in stratified mixtures with CxD < xst compared to the equiv-
flame, consider Fig. 24, which shows various shapes of the pdf of alent homogeneous mixture. However, in stratification with
the mixture fraction relative to the nominal lean and rich flam- CxD z xst, the turbulent flame speed may decrease over the homo-
mability limits. Assume that this pdf is homogeneous in space in geneous flame at xst. Turbulent flame speed correlations for strat-
the direction of flame propagation. If a flame propagates in mixture ified-charge flames are apparently not available. Turbulent
associated with a pdf of type S1 or S2, where the pdf resides fully stratified flames are an active field of study currently due to their
between xlean and xrich, we may speak of a stratified-charge premixed importance in a wide range of combustion applications and any
flame borrowing the terminology from spark-ignition engines. If advances made in their understanding and modelling are impor-
a flame propagates in mixture associated with a pdf of type E, then tant for the problem of ignition of non-premixed flames.
Renou et al. [260], in addition to studying the flame area
increase in their stratified flame, also examined the very early
stages of kernel formation in a flow of homogeneous isotropic
S2 S1 turbulence with uniform mean velocity and with a mixture fraction
Ρ(ξ) distribution characterised by a Gaussian pdf with mean CxD z xst
(xst ¼ 0.06 for their experiments) and r.m.s. between 0.010 and
0.014, varied by altering the fuel injection method. The low-x tail of
S3 the mixture fraction pdf just exceeded xlean and the high-x tail was
below xrich. Hence the fluid was mostly within the nominal flam-
mability limits. They observed some misfires (i.e. Pker < 1 in our
E terminology), but reported that for successful kernels, the level of
inhomogeneity did not affect the initial kernel radius.
ξlean ξst ξrich ξ

Fig. 24. Schematic of mixture fraction pdf’s relative to the nominal lean and rich
3.4.3. Turbulent edge flames
flammability limits, which can help distinguish between stratified-charge premixed Laminar flames propagating in regions with large mixture
flames (S1–S3) and non-premixed edge flames (E). fraction inhomogeneities may show three distinct reaction zones
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 87

[265] and are hence called triple flames. Under large enough scalar
dissipation that the lean and rich premixed wings merge, the flame
becomes an edge flame. There is a very large body of literature on
various aspects of these flames, reviewed by Buckmaster [34]. In
the context of ignition, we are interested in: (i) the propagation
speed; (ii) the maximum scalar dissipation rate for survival of an
edge flame; and (iii) how such flames propagate in turbulent flows.
Analysis (see Buckmaster [34] and Refs. [266,267] for an addi-
tional presentation of the main concepts and recent data) shows
that the propagation speed relative to the flow far ahead of the
triple point (defined as the point where the three fronts meet on
the stoichiometric isoline) is approximately given by Uedg-
1/2
e ¼ SL(xst)(r0/rb) with SL(xst) the laminar burning velocity of an
unstrained stoichiometric premixed flame, r0 the unburnt gas
density and rb the burnt gas density. This acceleration of the flame
is due to the divergent streamlines ahead of the edge, possibly first
visualised by Phillips [265]. When the edge is strained, its propa-
gation speed decreases and Uedge is additionally found to depend on
the fuel Lewis number [267,268]. The laminar edge flame is more
susceptible to straining out than an established flame. This is
because the heat flux from the propagating edge flame front is
Fig. 25. Isosurface of progress variable c ¼ 0.5 (red) superimposed on the stoichio-
higher than that of a continuous flame, as the flame front loses metric mixture fraction isosurface (blue) following sparking at the middle of the DNS
additional heat to the flow ahead of it. This can lead to extinction. A domain that contains isotropic decaying turbulence. [For interpretation of the refer-
theoretical prediction exists [269] that relates the ratio of the ences to colour in this figure legend, the reader is referred to the web version of this
maximum possible strain rate of the edge flame, Sed,ext, to the article.] Fuel is to the left of the xst isosurface, air to the right, straight interface initially.
Reprinted from Ref. [270], with permission from the American Institute of Physics.
extinction strain rate of the stable non-premixed flame, Sext: Sed,ext/
Sext ¼ e2b(1  3), where e ¼ 2.718, b ¼ E/RTb is the nondimensional
activation energy, E is the activation energy, R is the gas constant successful ignitions, it was found that increasing the turbulent
and Tb is the adiabatic flame temperature. As an estimate for intensity decreased the rate of expansion of the edge flame. This is
methane, E ¼ 160 kcal/mol and Tb ¼ 2226 K [5], thus the predicted contrary to our expectations from turbulent premixed flames,
ratio Sed,ext/Sext z 0.82. This prediction is close to the ratio between which overall move faster (up to the quenching point) with
the maximum velocity at which ignition can be achieved and the increasing turbulent intensity.
extinction velocity measured in turbulent non-premixed counter- The average edge flame speed (relative to fixed coordinates) was
flow flames [254] discussed previously, which suggests that part of not far from SL(xst)(r0/rb)1/2 [271]. Instantaneously, high positive
the reason why Pign < Pker may be the high strain rate on the edge and high negative values have also been observed. In addition, the
flame following the spark. edge flame propagation speed relative to the fluid also showed
Turbulent edge flames have been studied very little. The large scatter from various points of the stoichiometric isosurface,
extensive literature on turbulent lifted jet flames is of course very and the possibility of retreating fronts, rather than expanding, was
relevant, but few papers have measured directly the flow velocity at also observed. Misfires due to sparking far from stoichiometry have
the flame base. Various results are summarised by Lyons [20] who also been analysed [272]. The results so far obtained are for rela-
concludes that the mean flow speed approaching the flame is a few tively low turbulent Reynolds numbers and with one-step chem-
times greater than SL(xst) and that the streamline curvature at the istry. Further simulations with more realistic chemistry and for
edge slows down the fluid to the local edge speed that is approx- higher Reynolds numbers are necessary before we form a complete
imately SL(xst). The statistics of Uedge do not seem to have been picture on how turbulence affects the edge flame speed. Never-
measured yet directly in DNS or experiment. theless, the present data show an almost laminar-like mean edge
Recently, Chakraborty et al. [184,270–272] performed 3-D DNS flame speed along the stoichiometric contour following ignition. If
of fuel-air mixing layers with isotropic decaying turbulence and this result is confirmed from experiment and for higher Reynolds
with a localised energy source in the energy equation to mimic numbers, it implies that flame expansion in strongly turbulent
a spark. These simulations are the non-premixed ignition coun- flows following ignition is likely to rely more on turbulent disper-
terpart of well-established DNS of ignition in turbulent homoge- sion of the flame than on propagation of the flame edge relative to
neous mixtures [273]. Fig. 25 shows a progress variable isosurface the unburnt fluid along the stoichiometric contour.
(red) growing in all directions and intersecting the stoichiometric x Although there are no systematic data on turbulent flame
isosurface (blue). The intersection line forms the locus of triple speeds in stratified mixtures (Section 3.4.2), if we qualitatively
points of the turbulent edge flame. With weak or moderate u/SL, the consider those flames as a perturbation from turbulent fully-
flame expanded along the xst isosurface at a different speed than premixed combustion, there seems to be a fundamentally different
across x isosurfaces and hence the burnt region was not spherical behaviour concerning how turbulence affects the propagation of
(not even in the mean). With strong turbulence the kernel was stratified-charge premixed flames and non-premixed edge flames.
quenched and there was no propagation. The simulations also More research on both types of flames is necessary and in particular
allowed a detailed analysis of the displacement speed Sd, defined as on the conditions where they may quench. Such events contribute
the speed at which the fuel mass fraction isosurface moves with to the observed differences between the probability of creating
respect to an initially coincident material surface, at the triple a kernel but not igniting the whole flame (i.e. Pign < Pker) and are
points. Sd was found to depend on the local curvature of the flame, not well understood at present.
the fuel mass fraction gradient and the scalar dissipation [270] in
agreement with laminar flame studies (see the corresponding 3.4.4. Flame stabilisation
bibliography in Refs. [271,270]). Also, a very large scatter of values In non-premixed flame burners, the flame must be stabilised
for Sd was found along the flame’s edge. In general, even in after ignition. This is Phase 3 of the whole process. The usual means
88 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

of stabilisation of flames in high velocity gases is with a recircula- upstream flame evolution in igniting jets and the lift-off height. A
tion zone, which can be created by a bluff body or by strong swirl. later model [259] combining flamelet models for premixed and
Ignition of such flames has been studied, mostly to examine qual- non-premixed combustion and using a turbulent burning velocity
itatively the direction of flame propagation after sparking. The as a function of x also reproduced very well the lift-off heights.
available experiments with gaseous fuels comprise methane non- However, flame propagation was not tested. Veynante et al. [282]
premixed bluff-body flames [22], laser-induced ignition of propane modelled the flame propagation across and along x isosurfaces
[274] and prevapourized heptane [255] flames, laser-induced through an extended flame surface density model that included
ignition of cryogenic (H2/liquid oxygen) flames [275], and hydrogen premixed and non-premixed flamelets. At that time, there were no
bluff-body flames [276]. Older experiments with ignition of detailed experiments against which to compare, but the results
a turbulent premixed flame behind a V-gutter [277] have also been look plausible. For stratified-charge flames in engines, Drake and
reported. The common picture that emerges from these visual- Haworth [37] suggested that flamelet/PDF methods are suitable
isations is that a flame expands quickly in the downstream direc- (but no details were given), while Ref. [263] shows reasonable
tion following the flow, but also that the recirculation zone must be predictions of the pressure trace with a flamelet-based model for
fully ignited for successful overall flame establishment and stabi- premixed combustion and an Eddy Dissipation model for the
lisation. OH-PLIF images [22] show that as the flame propagates non-premixed combustion zone left behind the stratified-charge
into the recirculation zone a very ‘‘patchy’’ burnt region emerges, premixed flame (in the case studied, it is possible that the flame
which may lead to unburnt fluid and hence a recirculation zone corresponded to a mixture fraction pdf corresponding to curve S3
that is not as hot as it should be for good stability. Time traces of in Fig. 24 and hence this approach has merit). Many models have
chemiluminescence emission collected from the whole burner been developed [283–286], some of them derivatives from pre-
[274,275] show a very sharp peak associated with the spark, fol- mixed combustion theories and all showing a degree of success for
lowed by a less intense emission associated with the propagating their intended application. Bray et al. [287] discuss thoroughly
flame that consumes all the premixed fuel-air mixture in the some theoretical issues on the use of the progress variable in such
combustor volume, and ending with a much milder emission flames. The CMC model, which is conventionally used for non-
associated with the steady non-premixed flame. It is not easy to premixed combustion, has also been explored in the context of
distinguish failed ignition events at the stabilisation phase from turbulent ignition fronts by comparison with DNS [184] and it is
such traces due to the dominance of emission during Phases 1 and suggested that first-order closure of the chemical reaction rate term
2, but various trends on the flame expansion speed and the influ- is probably inadequate.
ence of the spark energy can be inferred. Large Eddy Simulation is possibly well suited to study spark
There is evidence from realistic gas turbine combustors ignition of non-premixed combustion. The mesh size may not be
[7,278,279] that successful ignition is associated with flame prop- too far (or even may be smaller than) the initial spark kernel and
agation upstream. Since the flow velocities are too high for direct hence the subsequent flame evolution could be tracked from its
propagation, the flame must arrive at the base of the recirculation very early stages. The unsteady nature of the flow will affect the
zone by convection by the recirculating flow. Despite the fact that flame evolution and, in principle, a measure of the probabilistic
the sparks used in gas turbines deposit large amounts of energy and nature of the process could be observed, for example, by per-
large kernels are virtually always created, successful flame estab- forming many LES runs for a given location of the spark which will
lishment does not always occur. The data from the simpler burners sample a different fluid mechanical field in every realisation.
suggest that this is due to the inability of flamelets to survive However, the way the initial spark can be modelled is an open
unquenched to the base of the recirculation zone [22]. Further work issue; experience from spark-ignition engine modelling may help
with planar imaging (for example with PLIF of a radical species or [283]. Finally, sub-grid closures for stratified or edge flames must
with Rayleigh scattering for temperature) during the ignition be included and recent results from the group at CERFACS with the
transient is necessary to elucidate the stabilisation process of LES code AVBP and the thickened-flame model extended to account
a grown flame in a combustor. for stratification and spray dispersion and evaporation [288,289]
are very encouraging. Very large scale parallel computations of the
3.5. Modelling approaches ignition sequence in helicopter combustors (Fig. 26) have been
performed and reasonable trends were shown, although unfortu-
Since ignition is a transient process and the flows we are nately there was no comparison with experiment. Effort should be
interested in are turbulent, the average behaviour must be under- directed towards properly validating such comprehensive models
stood as an ensemble over many ignition realisations. The average as they seem appropriate for treating all phases of the non-pre-
behaviour of the second and third phase of the process is easier to mixed flame ignition process.
understand, as it is conceptually similar to conventional premixed
flame propagation and hence suitable closure models in RANS or 3.6. Spark ignition of sprays
LES approaches should be sufficient. It is perhaps more difficult to
build into an average model the probabilistic nature of the kernel 3.6.1. Homogeneous dispersions
creation following a spark. The minimum ignition energy necessary to initiate a flame
In the context of predicting the ignition probability from a safety kernel in homogeneous dispersions of droplets in air, for a wide
point of view, Pker is considered equal to the flammability factor F variety of fuels, pressures, overall equivalence ratios, droplet sizes,
and hence the focus has been on the correct prediction of the and spark parameters has been examined by many authors, with
mixture fraction pdf that, for jets, must include intermittency, as a thorough review available [8]. Ballal and Lefebvre [238] proposed
suggested in Refs. [280,281]. The author is not aware of any other a correlation for MIE based on equating the time required for the
modelling effort to predict Pker. droplet-air mixture to evaporate and burn to the time required for
Following ignition, flame propagation in jets (without dis- the cold mixture to quench the kernel by diffusion (including
tinguishing between stratified-charge or turbulent edge flames) turbulent diffusion). This order-of-magnitude idea proves quite
has been modelled based on the G-equation and a laminar velocity, successful and results in predictions of minimum ignition energy
which is a function of the mixture fraction and which diminishes as that are close to experimental data covering a wide range of
the scalar dissipation increases [258]. This model seems quite well parameters. An alternative model, again based on characteristic
suited for turbulent edge flames, as it successfully captures the times and apparently produced independently, is also available
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 89

Fig. 26. Ignition sequence in an array of helicopter engine burners calculated with LES and the thickened-flame model. The colour denotes temperature (pink: low; yellow, red:
high). The incoming cold air coming from the swirlers and the dilution jets are evident in pink. [For interpretation of the references to colour in this figure legend, the reader is
referred to the web version of this article.] The thick white lines denote the reaction zones. Reproduced from Ref. [288], with Ref. [289] giving more details.

[290]. Both models are further discussed in the light of new data in gaseous case, consistent with an observed reduction in minimum
Refs. [291,292]. ignition energy under similar conditions. The effect of stretch,
Various fine details of the spark ignition of a droplet dispersion which is important for understanding the evolution of a spherical
have been revealed. Even with stagnant flow and a perfectly kernel following a spark, has been examined by Greenberg and
reproducible spark, there is a degree of stochasticity in the ignition Kalma [300] who concluded that the Lewis number plays a role in
resulting, for example, in an ignition probability that has a transi- the flame speed. Later, Greenberg [301] included velocity slip
tion from zero to unity over a range of spark energies. This is due to between the gas and the droplets and concluded that slip can cause
an inherent fluctuation of the actual fuel-air ratio in the sparked a significant alteration of the flame speed and even lead to
volume due to the random position of the droplets [293]. Ignition of extinction.
an overall lean droplet-air mixture may be easier than the corre-
sponding gaseous mixture at the same equivalence ratio due to the 3.6.2. Turbulent sprays
possibility of locally producing stoichiometric or rich equivalence Ignition in homogeneously-dispersed droplets in turbulent air
ratios that ignite with a smaller energy [8]. The Sauter mean flow has been studied experimentally by Ballal and Lefebvre and
diameter and the vapour partial pressure play a major role [294]. A a greater ignition energy was required in the presence of turbu-
multi-component fuel [295] and a polydisperse spray [292] can also lence [238], as expected from our insights from purely gaseous
alter the minimum ignition energy. The minimum ignition energy homogeneous turbulent flame ignition. Fig. 27 is taken from
with laser-induced sparks for droplet-air mixtures is discussed in Ref. [302] and shows temperature contours from a DNS of forced
Refs. [255,296]. ignition in turbulent droplet-air mixtures. In (b), the droplets are
Given that a kernel has been generated, the flame then propa- larger than in (a) and the ignition fails, as evidenced by the diffu-
gates in a mode that depends significantly on the droplet param- sion of the spark’s energy without causing self-sustaining
eters (volatility, overall equivalence ratio, size, temperature). In combustion. In these simulations, the droplets were treated as
case of very small droplets that vaporise quickly, a homogeneous point sources and hence the intra-droplet flame propagation
mixture will be attained at the leading edge of the propagating mechanism was not adequately resolved.
premixed flame. In the case of relatively big droplets, the flame may The propagating flame following ignition experiences locally
‘‘jump’’ from droplet to droplet with reaction zones surrounding a wide range of mixture fractions due to the random positions of
each droplet [3]. The speed of such flames has been measured and the droplets relative to the front that result in small-scale mixture
correlated with phenomenological models [297] and simulated fraction inhomogeneities. This has been demonstrated explicitly by
with detailed chemistry and a spray model [298]. In general, for DNS [303], where scatter plots showed a large range of mixture
smaller droplets the flame speed decreases over the purely gaseous fractions with high reaction rates, implying that the flame advances
case. Microgravity experiments have been examined by Niioka in a wide range of local stoichiometries. Modelling the evolution of
[299] who concluded that the highest propagation speed in heavy such flames is a very challenging task, as it includes concepts from
fuel droplets occurs when the inter-droplet spacing is approximately stratified-charge turbulent premixed flames and from turbulent
equal to the radius of individual droplet flames. In such a case and spray combustion. Very little is known on the turbulent flame
due to the possibility of establishing richer zones between the speed in fuel mists in turbulent flows, with the limited available
droplets, the overall flame speed may increase over the purely experimental data suggesting a faster speed with a decreasing
90 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

evolution following sparking under relight conditions [279,308]


and associated phenomenological modelling [308,309], and the
effects of the atomiser on ignition in a gas turbine combustor [278]
have been discussed. From such complicated flows it is not easy to
understand the phase where ignition failure occurred (kernel,
propagation, or stabilisation) and hence attention must be placed
to well-defined simplified experiments that can distinguish
between the various failure modes.

4. Conclusions and future research

This review has revealed various fundamental findings con-


cerning the autoignition and the forced ignition of turbulent non-
premixed flames. The turbulent flow affects these phenomena in
a way that is better understood for gaseous than for liquid fuels. A
summary of the main findings is attempted here, including
suggestions for future research.

4.1. Autoignition

Laminar flows under constant strain S, as in the counterflow


mixing layer between cold fuel and hot air (i.e. T1,0 < T2,0), will not
autoignite if the strain rate is above a critical value, Scrit. The criti-
cality can also be expressed in terms of a critical scalar dissipation
Ncrit. Experimental data often report the critical air temperature,
Tcrit, below which autoignition does not happen for the given S. It is
found that virtually always Tcrit increases with S. Laminar parabolic
flows, such as an unsteady mixing layer, always autoignite, but at
a time sign that increases with increasing N. Hence the initial
condition of the interface between fuel and air will affect the
autoignition time. In unsteady mixing layers with constant strain,
numerical simulations show that sign increases with N/Ncrit at first
mildly. As N/Ncrit approaches unity, sign rapidly approaches infinity,
i.e. no autoignition, due to too high scalar dissipation. Differential
diffusion can alter the calculated sign and quantification of Ncrit
through numerical simulations or experiments is important.
Fig. 27. Slice through the DNS computational domain showing temperature (blue: cold
The physical location of autoignition can be identified through
reactants temperature; red: stoichiometric adiabatic flame temperature) following
ignition from a power source inside the thick dashed circle in a droplet-air mist. [For its location in mixture fraction space and the (known) mixture
interpretation of the references to colour in this figure legend, the reader is referred to fraction spatial distribution. The location of autoignition in mixture
the web version of this article.] Initially, a homogeneous droplet dispersion corre- fraction space is denoted as the most reactive mixture fraction xMR,
sponding to an overall equivalence ratio of 2 was placed in a y–z slab of thickness which can be quantified by various methods, all giving results that
about 1/2 of the length of domain in the x-direction. No vapour existed initially.
Droplets within slice are represented by dots and unbroken lines represent contours of
are reasonably similar. It is important to quantify xMR with some
mixture fraction: xst by thick white lines; 4xst/3 by thin red lines. (a) sevap/(D/S2L ) ¼ 1.95; method before considering the autoignition of a turbulent non-
(b) sevap/(D/S2L ) ¼ 2.83 (D: molecular diffusivity, SL: stoichiometric laminar burning premixed system. Note that xMR is not an intrinsic property of the
velocity). The longer evaporation time in (b) causes a smaller generated mixture fuel as it depends on the operating conditions. In simulations, it
fraction and misfire. Reproduced from Ref. [302], with permission from the Combus-
may also depend on the chemical mechanism chosen. It is also
tion Institute.
important to note that the mixture fraction at which the reaction
rate peaks may shift during the induction time, and that this shift
Sauter mean diameter and for more volatile fuels [304,305]. More depends on the fuel and the scalar dissipation. This shift may also
experimental work in this area is necessary. reflect fundamental changes in the chemical structure of the
In turbulent sprays, such as those in realistic combustion system, e.g. switching from a chain branching to a thermal explo-
devices, the number density of droplets is a strong function of sion mode.
space. Most of our knowledge on turbulent combustion of such In a turbulent non-premixed flow with T1,0 < T2,0, extensive
sprays comes from established rather than propagating flames in evidence from DNS (using both 2-D and 3-D codes, with simple and
turbulent droplet-air flows. Direct Numerical Simulations covering complex chemistry, various fuels and initial fuel-air distributions),
a wide range of evaporation, flame, and turbulent timescales and shows that the mixture fraction that first autoignites is close to the
droplet spacing and diameters, have been used to examine flame xMR defined from stagnant mixture or laminar flow analyses.
topologies in jet-like flows [306]. DNS can help the development of However, in the turbulent flow, not all regions with x ¼ xMR auto-
models for flame propagation in droplet dispersions and the ignite simultaneously, giving hence a certain ‘‘spotiness’’ to the
particular effects of a non-uniform droplet number density on process. Those regions that have low scalar dissipation autoignite
ignition and on flame propagation must be explored. earlier than those with high N. Such regions can be cores of vortices
In contrast to our fundamental knowledge of turbulent spray or locations of the xMR isoline that are concave to the air; the effect
flame ignition that is hardly adequate, practical studies of turbulent of differential diffusion becomes important here. Hence, the fluc-
spray ignition seem to have been more extensive due to the tuations of the scalar dissipation are important for autoignition,
importance of successfully igniting spray flames in aviation jet and this importance is greatest if N is high and close to Ncrit. The
engines. Hence, the type of ignitor [7,307], visualisation of flame history of the conditional NjxMR is important: autoignition spots are
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 91

those that, in general, have experienced low NjxMR. The effects of process. Since the main emphasis of such work would be the effects
initial thickness of the fuel-air interface, initial turbulent velocity of turbulence on the process rather than the chemistry, it is
and initial turbulent lengthscale seem to affect the emergence of emphasized that proper measurements of the turbulent velocity
the first autoignition site through their action on the conditional and mixture fraction fields are performed. There is no explicit
scalar dissipation. The spatial statistics of the sites are determined experimental validation yet of the DNS finding that autoignition
by the spatial spreading of the low NjxMR regions; very little is occurs at low N; it would be challenging but very insightful to
known about this from either experiment or DNS. The statistics of attempt this in experiment.
sign from individual realisations of the DNS seem to follow a very On the modelling side, we must develop further CMC using
narrow distribution, but more data focused on this aspect are second-order and PDF methods using joint velocity-frequency-
necessary. Due to the large importance of the fluctuations of the scalar approaches, both of which are likely to include in a better
scalar dissipation rate, modelling approaches best suited to this manner the fluctuations of mixing rate that seem to drive the
problem are methods that include the fluctuations of the mixing spotiness of autoignition kernels. Double-conditioned CMC must
rate. The magnitude of sign in the turbulent flow as compared to sref also be tried for these flows. The new flamelet methods based on
is important to quantify. In all cases where this quantification has two (or more) mixture fractions are also interesting to pursue. LES
been made, it turns out that sign/sref > 1, which implies that we can efforts seem very promising.
understand sref as the minimum possible autoignition time in In simulations of turbulent non-premixed autoignition prob-
a non-premixed situation even in the presence of turbulence. lems, we are trying to reproduce quite fine trends on how turbu-
Most of the work done so far, even when complex chemistry has lence affects the autoignition process, and hence the chemical
been used, covers regions where the fuel behaves in an approxi- mechanism used must be very reliable. Mechanisms validated
mately Arrhenius manner, which implies that autoignition is for both low- and high-temperature chemistry are necessary,
decelerated when the temperature decreases. This may not be the since the overall ignition process involves flame propagation. For
case with hydrocarbons under some conditions that show the a good assessment of whether a turbulent combustion model
Negative Temperature Coefficient regime, where autoignition time reproduces experimentally-observed autoignition lengths or times
increases as temperature increases. In this case, it is not evident in a turbulent non-premixed flow, the mechanism must probably
that an increased scalar dissipation will delay autoignition and the be in a position to reproduce very accurately the autoignition time
xMR concept has not been explored. Simulations and experiments of homogeneous mixtures in the conditions of the experiment and
targeted at the effects of scalar dissipation and its fluctuations on not necessarily to perform well in a wide range of pressures or
this regime of operation are needed to clarify these issues. When initial temperatures.
reactions zones are wide, the additional effect of large scalar
dissipation rate variations across the reaction zone must also be 4.2. Spark ignition
considered. The effect of pressure, and its influence on the domi-
nant chemical pathways, has not been adequately studied for Spark ignition of homogeneous stagnant mixtures can fail if the
turbulent non-premixed systems. energy delivered is not sufficient to drive a flame kernel above
Following localised autoignition, limited experimental and a certain critical radius that is proportional to the laminar flame
numerical evidence shows that flamelets grow around the auto- thickness. In the presence of turbulence, the additional strain rate
ignition kernel at speeds close in magnitude to the laminar burning can quench the kernel in its early stages and hence lead to misfire.
velocity of a stoichiometric mixture at an unburnt temperature Spark ignition of laminar non-premixed flames has been very little
T0(xst). These flamelets may collide with separate flamelets studied. The very limited analytical and numerical results that exist
emanating from neighbouring spots, igniting hence the whole show that the ignition may fail if the energy is not above a given
flame. If fresh mixture is continuously flowing, this distributed threshold, which now depends on the strain rate and the location of
process may result in a well-defined stabilisation region that the spark relative to the (nominally) flammable region. Locally,
provides overall flame initiation. Such situations have been simu- forced ignition even at stoichiometry may fail if the strain rate is too
lated with detailed-chemistry DNS for jets of hydrogen in hot high. Long-range effects have been observed, where the diffusion or
oxidizer. These simulations show local fluctuations of the ignition convection of heat from the spark may eventually reach the flam-
point and demonstrate that autoignition is responsible for the mable region and ignite the flame. Further experimental and
stabilisation. theoretical investigations of laminar non-premixed flame spark
Very few experimental investigations have focused on turbulent ignition are necessary.
non-premixed autoignition. With continuous flow, the configura- In turbulent flows, ignition has a stochastic nature that depends
tions studied so far are the turbulent counterflow between fuel and not only on whether the spark samples flammable mixture, but also
electrically-heated air, where the critical air temperature has been on the local strain that may quench the kernel. Hence, we distin-
quantified, the jet in vitiated air, and a fuel jet or plume in fast co- guish the probability of initiating a kernel, Pker, the probability of
flowing enclosed heated air. The open flow provides good optical establishing the whole flame, Pign, and the probability of finding
access, facilitating therefore point and planar diagnostics that have flammable mixture fractions that is called the flammability factor F
visualised pre-ignition regions, autoignition spots, and the ensuing (Eq. (5)). Experiments in turbulent jets, counterflow and bluff-body
flamelets. Point measurements of mixture fraction are consistent stabilised flames show that Pign < Pker in many places in the flow
with the view from DNS that ignition first happens at the most depending on the success of flame propagation and overall flame
reactive mixture fraction, although a detailed comparison has not stabilisation and that Pker is not necessarily equal to F due to local
been attempted. The enclosed flow allows a more direct quantifi- quenching at the early stages of the flame. More experiments and
cation of the background turbulence on the autoignition, and simultaneous measurements of mixture fraction and scalar dissi-
a retardation of autoignition due to intense turbulence has been pation at the spark location are necessary to understand better the
reported. With transient flows, jets injected in high-pressure, reasons why F and Pker may differ.
diesel-like environments have been used with natural gas and The nature of flame propagation in mixtures with large mixture
DME, but few other fuels seem to have been studied. fraction fluctuations must be explored. For turbulent edge flames
A lot remains to be done experimentally using more fuels and (flames propagating mostly along the stoichiometric contour in
fast laser diagnostics in order to identify the statistics of the auto- flows with wide pdf’s of the mixture fraction), DNS shows that their
ignition sites and to analyse better the overall flame growth propagation speed is small and on average decreases with
92 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

increasing turbulent intensity and hence ignition of the overall [10] Liñán A, Crespo A. Asymptotic analysis of unsteady diffusion flames for large
activation energies. Combustion Science and Technology 1976;14:95–117.
flame depends mostly on convection and turbulent diffusion.
[11] Mastorakos E, Baritaud TA, Poinsot TJ. Numerical simulations of autoignition
Turbulent stratified-charge flames (i.e. flames propagating in in turbulent mixing flows. Combustion and Flame 1997;109:198–223.
mixture fraction pdf’s not exceeding substantially the nominal [12] Dopazo C, O’Brien EE. An approach to the autoignition of a turbulent mixture.
flammability limits) are probably faster and their response to Acta Astronautica 1974;1:1239–66.
[13] Johansson B. Homogeneous charge compression ignition: the future of IC
turbulence is more akin to premixed turbulent flames. Both types of engines? International Journal of Vehicle Design 2007;44:1–19.
flames may be present in the transient ignition process of non- [14] Sankaran R, Im HG, Hawkes ER, Chen JH. The effects of non-uniform
premixed combustion and must be understood further. Modelling temperature distribution on the ignition of a lean homogeneous hydrogen-
air mixture. Proceedings of the Combustion Institute 2005;30:875–82.
of forced ignition of turbulent non-premixed flames is very limited. [15] Hawkes ER, Sankaran R, Pebay PP, Chen JH. Direct numerical simulation of
Large Eddy Simulations seem very suitable for capturing some of ignition front propagation in a constant volume with temperature inhomo-
the variability of behaviour observed. geneities II. Parametric study. Combustion and Flame 2006;145:145–59.
[16] Chen JH, Hawkes ER, Sankaran R, Mason SD, Im HG. Direct numerical
Additional candidate topics for future research are: (i) turbulent simulation of ignition front propagation in a constant volume with temper-
edge flame propagation in inhomogeneous mixtures; (ii) the ature inhomogeneities I. Fundamental analysis and diagnostics. Combustion
transient stabilisation process of recirculating flames; (iii) the and Flame 2006;145:128–44.
[17] Orrin JE, Vince IM, Weinberg FJ. A study of plasma jet ignition mechanisms.
nature of the flame generation process at very short timescales, i.e. Proceedings of the Combustion Institute 1981;18:1755–65.
before any appreciable propagation, by sparking in laminar and [18] Takita K, Moriwaki A, Kitagawa T, Masuya G. Ignition and flame-holding of H2
turbulent inhomogeneous mixtures. and CH4 in high temperature airflow by a plasma torch. Combustion and
Flame 2003;132:679–89.
[19] Bilger RW, Pope SB, Bray KNC, Driscoll JF. Paradigms in turbulent combustion
4.3. Spray ignition research. Proceedings of the Combustion Institute 2005;30:21–42.
[20] Lyons KM. Toward an understanding of the stabilization mechanisms of lifted
turbulent jet flames: experiments. Progress in Energy and Combustion
Autoignition and forced ignition of sprays in stagnant or laminar Science 2007;33:211–31.
flows has been extensively studied and reviewed elsewhere, but [21] Masri AR, Kalt PAM, Barlow RS. The compositional structure of swirl-stabi-
the effects of turbulence on these processes are still not clear. The lised turbulent nonpremixed flames. Combustion and Flame 2004;137:1–37.
[22] Ahmed SF, Balachandran R, Marchione T, Mastorakos E. Spark ignition of
effects of evaporation on mixing seems a very important issue for turbulent non-premixed bluff-body flames. Combustion and Flame
extending our knowledge and for increasing our predictive capa- 2007;151:366–85.
bilities from gaseous to spray autoignition. Turbulent flame prop- [23] Birch AD, Brown DR, Dodson MG, Thomas JR. Studies of flammability in
turbulent flows using laser Raman spectroscopy. Proceedings of the
agation in sprays, in particular when the droplet number density Combustion Institute 1978;17:307–14.
has strong spatial variations, and the eventual flame stabilisation [24] Birch AD, Brown DR, Dodson MG, Thomas JR. Turbulent concentration field of
seem the crucial topics concerning spark ignition in direct-injection a methane jet. Journal of Fluid Mechanics 1978;88:431–49.
[25] Akindele OO, Bradley D, Mak PW, McMahon M. Spark ignition of turbulent
engines and gas turbines. The development of high repetition gases. Combustion and Flame 1982;47:129–55.
diagnostics for sprays and the development of modelling [26] Simmie JM. Detailed chemical kinetic models for the combustion of hydro-
approaches capturing both premixed and non-premixed reaction carbon fuels. Progress in Energy and Combustion Science 2003;29:599–634.
[27] Miller JA, Pilling MJ, Troe J. Unravelling combustion mechanisms through
zones, of varying degree of completion, in gaseous and spray a quantitative understanding of elementary reactions. Proceedings of the
combustion are necessary. Combustion Institute 2005;30:43–88.
[28] Sung C-J, Law CK. Fundamental combustion properties of H2/CO mixtures:
ignition and flame propagation at elevated pressures. Combustion Science
Acknowledgements and Technology 2008;180:1097–116.
[29] Buckmaster J, Clavin P, Liñán A, Matalon M, Peters N, Shivashinsky G, et al.
Combustion theory and modelling. Proceedings of the Combustion Institute
The author wishes to acknowledge the contribution of Profs. R.
2005;30:1–19.
W. Bilger and T. J. Poinsot through useful discussions on various [30] Sazhin SS. Advanced models of fuel droplet heating and evaporation. Prog-
aspects of ignition over many years. Collaborations with Dr. C. ress in Energy and Combustion Science 2006;32:162–214.
Frouzakis at ETH, Prof. A. Masri at Sydney, and Drs. R. Eggels, P. [31] Cavaliere A, de Joannon M. Mild combustion. Progress in Energy and
Combustion Science 2004;30:329–66.
Schober, and J. Moran from Rolls-Royce have been very productive. [32] Pellett GL, Bruno C, Chinitz W. Review of air vitiation effects on scramjet
At Cambridge, the work of Drs. S. F. Ahmed, N. Chakraborty, C. N. ignition and flameholding combustion processes. AIAA Paper 2002. 2002-
Markides, G. De Paola, and E. S. Richardson has shaped this review; 3880.
[33] Curran ET. Scramjet engines: the first forty years. Journal of Propulsion and
thanks to them also for their useful comments. I am grateful to Power 2001;17:1138–48.
Profs. A. Masri and M. Fairweather for their suggestions on a draft of [34] Buckmaster J. Edge-flames. Progress in Energy and Combustion Science
this paper and for their encouragement. Financial assistance has 2002;28:435–75.
[35] Dale JD, Checkel MD, Smy PR. Application of high energy ignition systems to
been provided by the European Commission, the Engineering and engines. Progress in Energy and Combustion Science 1997;23:379–98.
Physical Sciences Research Council, Ford Research Centre Aachen [36] Mellor AM. Semi-empirical correlations for gas turbine emissions, ignition, and
and Rolls-Royce Group plc. flame stabilization. Progress in Energy and Combustion Science 1980;6:347–
58.
[37] Drake MC, Haworth DC. Advanced gasoline engine development using optical
References diagnostics and numerical modeling. Proceedings of the Combustion Insti-
tute 2007;31:99–124.
[38] Boehman AL, Le Corre O. Combustion of syngas in internal combustion
[1] Spalding DB. Combustion and mass transfer. Oxford: Pergamon Press; 1979.
engines. Combustion Science and Technology 2008;180:1193–206.
[2] Law CK. Combustion physics. Cambridge: Cambridge University Press; 2006.
[39] Lieuwen T, McDonell V, Santavicca D, Sattelmayer T. Burner development and
[3] Williams FA. Combustion theory. Menlo Park, California: The Benjamin/
operability issues associated with steady flowing syngas fired combustors.
Cummings Publishing Company; 1985.
Combustion Science and Technology 2008;180:1169–92.
[4] Glassman I. Combustion. 2nd ed. Academic Press; 1987.
[40] Dryer FL, Chaos M, Zhao Z, Stein JN, Alpert JY, Homer CJ. Spontaneous ignition
[5] Turns SR. An introduction to combustion: concepts and applications. London:
of pressurized releases of hydrogen and natural gas into air. Combustion
McGraw Hill; 2000.
Science and Technology 2007;179:663–94.
[6] Lewis B, von Elbe G. Combustion, flames and explosions of gases. , London:
[41] Astbury GR, Hawksworth SJ. Spontaneous ignition of hydrogen leaks:
Harcourt Brace Jovanovich Publishers; 1987.
a review of postulated mechanisms. International Journal of Hydrogen
[7] Lefebvre AH. Gas turbine combustion. 2nd ed. London: Taylor and Francis;
Energy 2007;32:2178–85.
1998.
[42] Hirano T. Combustion science for safety. Proceedings of the Combustion
[8] Aggarwal SK. A review of spray ignition phenomena: present status and
Institute 2002;29:167–80.
future research. Progress in Energy and Combustion Science 1998;24:565–
[43] Proust Ch. A few fundamental aspects about ignition and flame propaga-
600.
tion in dust clouds. Journal of Loss Prevention in the Process Industries
[9] Liñán A. The asymptotic structure of counterflow diffusion flames for large
2006;19:104–20.
activation energies. Acta Astronautica 1974;1:1007–39.
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 93

[44] Veynante D, Vervisch L. Turbulent combustion modeling. Progress in Energy [75] Han B, Sung CJ, Nishioka M. Effects of vitiated air on hydrogen ignition in
and Combustion Science 2002;28:193–266. a high-speed laminar mixing layer. Combustion Science and Technology
[45] Cant RS, Mastorakos E. An introduction to turbulent reacting flows. London: 2004;176:305–30.
Imperial College Press; 2008. [76] Takita K, Abe N, Mausya G, Ju Y. Ignition enhancement by addition of NO and
[46] Pitsch H. Large-eddy simulation of turbulent combustion. Annual Review of NO2 from a N2/O2 plasma torch in a supersonic flow. Proceedings of the
Fluid Mechanics 2006;38:453–82. Combustion Institute 2007;31:2489–96.
[47] Peters N. Turbulent combustion. Cambridge University Press; 2000. [77] Tanaka S, Ayala F, Keck JC, Heywood JB. Two-stage ignition in HCCI
[48] Bilger RW. Some aspects of scalar dissipation. Flow, Turbulence and combustion and HCCI control by fuels and additives. Combustion and Flame
Combustion 2004;72:93–114. 2003;132:219–39.
[49] Seiser R, Seshardi K, Piskernik E, Liñán A. Ignition in the viscous layer [78] Thevenin D, Candel S. Effect of variable strain on the dynamics of diffusion
between counterflowing streams: asymptotic theory with comparison to flame ignition. Combustion Science and Technology 1993;91:73–94.
experiments. Combustion and Flame 2000;122:339–49. [79] Thevenin D, Candel S. Ignition dynamics of a diffusion flame rolled up in
[50] Sung CJ, Law CK. Ignition of oscillatory counterflowing nonpremixed a vortex. Physics of Fluids 1995;7:434–45.
hydrogen against heated air. Combustion Science and Technology 1997;129: [80] Yetter RA, Dryer FL, Rabitz H. A comprehensive reaction-mechanism for
347–70. carbon-monoxide hydrogen oxygen kinetics. Combustion Science and Tech-
[51] Im HG, Raja LL, Kee RJ, Petzold LR. A numerical study of transient ignition in nology 1991;79:97–128.
a counterflow nonpremixed methane-air flame using adaptive time inte- [81] Hilbert R, Thevenin D. Autoignition of turbulent non-premixed flames
gration. Combustion Science and Technology 2000;158:341–63. investigated using direct numerical simulations. Combustion and Flame
[52] Fotache CG, Kreutz TG, Zhu DL, Law CK. An experimental study of ignition in 2002;128:22–37.
nonpremixed counterflowing hydrogen versus heated air. Combustion [82] Kim S-K, Yu Y, Ahn J, Kim Y-M. Numerical investigation of the autoignition of
Science and Technology 1995;109:373–93. turbulent gaseous jets in a high-pressure environment using the multiple-
[53] Trujillo JYD, Kreutz TG, Law CK. Ignition in a counterflowing non-premixed RIF model. Fuel 2004;83:375–86.
CO/H2-air system. Combustion Science and Technology 1997;127:1–27. [83] Gopalakrishnan V, Abraham J. Effects of multicomponent diffusion on pre-
[54] Fotache CG, Tan Y, Sung CJ, Law CK. Ignition of CO/H2/N2 versus heated air in dicted ignition characteristics of an n-heptane diffusion flame. Combustion
counterflow: experimental and modelling results. Combustion and Flame and Flame 2004;136:557–66.
2000;120:417–26. [84] Wright YM, De Paola G, Boulouchos K, Mastorakos E. Simulations of spray
[55] Fotache CG, Kreutz TG, Law CK. Ignition of counterflowing methane versus autoignition and flame establishment with two-dimensional CMC.
heated air under reduced and elevated pressures. Combustion and Flame Combustion and Flame 2005;143:402–19.
1997;108:442–70. [85] Massias A, Diamantis D, Mastorakos E, Goussis DA. An algorithm for the
[56] Fotache CG, Kreutz TG, Law CK. Ignition of hydrogen-enriched methane by construction of global reduced mechanisms with CSP data. Combustion and
heated air. Combustion and Flame 1997;110:429–40. Flame 1999;117:685–708.
[57] Fotache CG, Wang H, Law CK. Ignition of ethane, propane, and butane in [86] Lovas T, Amneus P, Mauss F, Mastorakos E. Comparison of automatic reduc-
counterflow jets of cold fuel versus hot air under variable pressures. tion procedures for ignition chemistry. Proceedings of the Combustion
Combustion and Flame 1999;117:777–94. Institute 2002;29:1387–93.
[58] Jomaas G, Zheng XL, Zhu DL, Law CK. Experimental determination of coun- [87] Ju Y. Asymptotic and numerical analyses of ignition of a fuel jet in a super-
terflow ignition temperatures and laminar flame speeds of C2–C3 hydro- sonic airstream. Combustion Science and Technology 1995;108:47–65.
carbons at atmospheric and elevated pressures. Proceedings of the [88] Knikker R, Dauptain A, Cuenot B, Poinsot T. Comparison of computational
Combustion Institute 2005;30:193–200. methodologies for ignition of diffusion layers. Combustion Science and
[59] Kortschik C, Honnet S, Peters N. Influence of curvature on the onset of Technology 2003;175:1783–806.
autoignition in a corrugated counterflow mixing field. Combustion and [89] Liñán A, Williams FA. Autoignition of nonuniform mixtures in chambers of
Flame 2005;142:140–52. variable volume. Combustion Science and Technology 1995;105:245–63.
[60] Holley AT, Dong Y, Andac MG, Egolfopoulos FN, Edwards T. Ignition and [90] Sánchez AL. Nonpremixed spontaneous ignition in the laminar wake of
extinction of non-premixed flames of single-component liquid hydrocar- a splitter plate. Physics of Fluids 1997;9:2032–44.
bons, jet fuels, and their surrogates. Proceedings of the Combustion Institute [91] Dooley DA. Combustion in laminar mixing regions and boundary layers. PhD
2007;31:1205–13. thesis. California Institute of Technology; 1956.
[61] Egolfopoulos FN, Dimotakis PE. Non-premixed hydrocarbon ignition at high [92] Zheng XL, Yuan J, Law CK. Nonpremixed ignition of H2/air in a mixing layer
strain rates. Proceedings of the Combustion Institute 1998;27:641–8. with a vortex. Proceedings of the Combustion Institute 2005;30:415–21.
[62] Zheng XL, Lu TF, Law CK, Westbrook CK, Curran HJ. Experimental and [93] Mastorakos E, Pires Da Cruz A, Baritaud TA, Poinsot TJ. A model for the effects
computational study of nonpremixed ignition of dimethyl ether in counter- of mixing on the autoignition of turbulent flows. Combustion Science and
flow. Proceedings of the Combustion Institute 2005;30:1101–9. Technology 1997;125:243–82.
[63] Kreutz TG, Nishioka M, Law CK. The role of kinetic versus thermal feedback in [94] Sreedhara S, Lakshmisha KN. Direct numerical simulation of autoignition in
nonpremixed ignition of hydrogen versus heated air. Combustion and Flame a non-premixed, turbulent medium. Proceedings of the Combustion Institute
1994;99:758–66. 2000;28:25–34.
[64] Balakrishnan G, Smooke MD, Williams FA. A numerical investigation of [95] Cao S, Echekki T. Autoignition in nonhomogeneous mixtures: conditional
extinction and ignition limits in laminar nonpremixed counterflowing statistics and implications for modeling. Combustion and Flame 2007;151:
hydrogen-air streams for both elementary and reduced chemistry. 120–41.
Combustion and Flame 1995;102:329–40. [96] Im HG, Chen JH, Law CK. Ignition of hydrogen-air mixing layer in turbulent
[65] Liu S, Hewson JC, Chen JH, Pitsch H. Effects of strain rate on high-pressure flows. Proceedings of the Combustion Institute 1998;27:1047–56.
nonpremixed n-heptane autoignition in counterflow. Combustion and Flame [97] Echekki T, Chen JH. Direct numerical simulation of autoignition in non-
2004;137:320–39. homogeneous hydrogen-air mixtures. Combustion and Flame 2003;
[66] Seshardi K, Peters N, Paczko G. Rate-ratio asymptotic analysis of autoignition 134:169–91.
of n-heptane in laminar nonpremixed flows. Combustion and Flame [98] Viggiano A, Magi V. A 2-D investigation of n-heptane autoignition by means
2006;146:131–41. of direct numerical simulation. Combustion and Flame 2004;137:432–43.
[67] Andac MG, Egolfopoulos FN. Diffusion and kinetics effects on the ignition of [99] Sreedhara S, Lakshmisha KN. Autoignition in a non-premixed medium: DNS
premixed and non-premixed flames. Proceedings of the Combustion Institute studies on the effects of three-dimensional turbulence. Proceedings of the
2007;31:1165–72. Combustion Institute 2002;29:2051–9.
[68] Liu S, Hewson JC, Chen JH. Nonpremixed n-heptane autoignition in unsteady [100] van Kalmthout E. PhD thesis. Ecole Centrale de Paris; 1996.
counterflow. Combustion and Flame 2006;145:730–9. [101] Yoo CS, Chen JH, Sankaran R. 3-D direct numerical simulation of
[69] Mason SD, Chen JH, Im HG. Effects of unsteady scalar dissipation rate on turbulent lifted hydrogen/air jet flame in heated coflow. Submitted for
ignition of non-premixed hydrogen/air mixtures in counterflow. Proceedings publication.
of the Combustion Institute 2002;29:1629–36. [102] Kerkemeier S, Frouzakis CE, Tomboulides AG, Mastorakos E, Boulouchos K,
[70] Langille JA, Dong Y, Andac MG, Egolfopoulos FN, Tsotsis TT. Non-premixed Autoignition of a diluted hydrogen jet in a heated 2-D turbulent air flow.
ignition by vitiated air in counterflow configurations. Combustion Science Conference on DNS and LES of reacting flows, 22–24 October, Technische
and Technology 2006;178:635–53. Universiteit Eindhoven, The Netherlands; 2008.
[71] Mastorakos E, Taylor AMKP, Whitelaw JH. Extinction of turbulent counter- [103] Markides CN, Mastorakos E. An experimental study of hydrogen autoignition
flow flames with reactants diluted by hot products. Combustion and Flame in a turbulent co-flow of heated air. Proceedings of the Combustion Institute
1995;102:101–14. 2005;30:883–91.
[72] De Joannon M, Matarazzo A, Sabia P, Cavaliere A. Mild combustion in [104] Li J, Zhao ZW, Kazakov A, Dryer FL. An updated comprehensive kinetic model
homogeneous charge diffusion ignition (HCDI) regime. Proceedings of the of hydrogen combustion. International Journal of Chemical Kinetics
Combustion Institute 2007;31:3409–16. 2004;36:566–75.
[73] Tan Y, Fotache CG, Law CK. Effects of NO on the ignition of hydrogen and [105] Del Alamo G, Williams FA, Sanchez AL. Hydrogen-oxygen induction times
hydrocarbons by heated counterflowing air. Combustion and Flame above crossover temperatures. Combustion Science and Technology
1999;119:346–55. 2004;176:1599–626.
[74] Egolfopoulos FN, Dimotakis PE. Effects of additives on the non-premixed [106] O’Conaire M, Curran HJ, Simmie JM, Pitz WJ, Westbrook CK. A comprehensive
ignition of ethylene in air. Combustion Science and Technology 2000;156: modeling study of hydrogen oxidation. International Journal of Chemical
173–99. Kinetics 2004;36:603–22.
94 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

[107] Domingo P, Vervisch L. Triple flames and partially premixed combustion in [137] Gordon RL, Masri AR, Pope SB, Goldin GM. A numerical study of auto-ignition
autoignition of non-premixed turbulent mixtures. Proceedings of the in turbulent lifted flames issuing into a vitiated co-flow. Combustion Theory
Combustion Institute 1996;26:233–40. and Modelling 2007;11:351–76.
[108] Aleiferis PG, Charalambides AG, Hardalupas Y, Taylor AMKP, Urata Y. [138] Terry SD, Lyons KM. Low Reynolds number turbulent lifted flames in high co-
Modelling and experiments of HCCI engine combustion with charge strati- flow. Combustion Science and Technology 2005;177:2091–112.
fication and internal EGR. SAE Paper 2005. 05FFL-189. [139] Siebers D, Higgins B. Flame lift-off on direct-injection diesel sprays under
[109] Hasegawa T, Arai A, Kadowaki S, Yamaguchi S. Autoignition of a turbulent quiescent conditions. SAE Paper 2001. 2001-01-0530.
premixed gas. Combustion Science and Technology 1992;84:1–13. [140] Medwell PR, Kalt PAM, Dally BB. Imaging of diluted turbulent ethylene flames
[110] Cabra T, Myhrvold JY, Chen RW, Dibble AN, Karpetis, Barlow RS. Simultaneous stabilised on a jet in hot coflow (JHC) burner. Combustion and Flame
laser Raman–Rayleigh-LIF measurements and numerical modeling results of 2008;152:100–13.
a lifted turbulent H2/N2 jet flame in a vitiated coflow. Proceedings of the [141] Singh S, Reitz RD, Musculus MPB. Comparison of the characteristic time
Combustion Institute 2002;29:1881–8. (CTC), representative interactive flamelet (RIF), and direct integration with
[111] Cabra R, Chen J-Y, Dibble RW, Karpetis AN, Barlow RS. Lifted methane–air jet detailed chemistry combustion models against optical diagnostic data for
flames in a vitiated coflow. Combustion and Flame 2005;143:491–506. multi-mode combustion in a heavy-duty DI diesel engine. SAE Paper 2006.
[112] Wu Z, Masri AR, Bilger RW. An experimental investigation of the turbulence 2006-01-0055.
structure of a lifted H2/N2 jet flame in a vitiated co-flow. Flow, Turbulence [142] Ertesvåg IS, Magnussen BF. The eddy dissipation turbulence energy cascade
and Combustion 2006;76:61–81. model. Combustion Science and Technology 2000;159:213–35.
[113] Vedula P, Yeung PK, Fox RO. Dynamics of scalar dissipation in isotropic [143] Hong S, Wooldridge MS, Assanis DN. Modeling of chemical and mixing
turbulence: a numerical and modeling study. Journal of Fluid Mechanics effects of methane autoignition under direct-injection, stratified charged
2001;433:29–60. conditions. Proceedings of the Combustion Institute 2002;29:711–8.
[114] Sreedhara S, Lakshmisha KN. Assessment of conditional moment closure [144] Myhrvold T, Ertesvåg IS, Gran IR, Cabra R, Chen J-Y. A numerical investigation
models of turbulent autoignition using DNS data. Proceedings of the of a lifted H2/N2 turbulent jet flame in a vitiated coflow. Combustion Science
Combustion Institute 2002;29:2069–77. and Technology 2006;178:1001–30.
[115] Sreenivasan K. Possible effects of small-scale intermittency in turbulent [145] Lakshmisha KN, Zhang Y, Rogg B, Bray KNC. Modelling auto-ignition in
reacting flows. Flow, Turbulence and Combustion 2004;72:115–31. a turbulent medium. Proceedings of the Combustion Institute 1992;24:
[116] Gaydon AG, Moore NPW, Simonson JR. Chemical and spectroscopic studies of 421–8.
blue flames in the auto-ignition of methane. Proceedings of the Royal Society [146] Lakshmisha KN, Rogg B, Bray KNC. On the regimes of auto-ignition in
of London, Series A 1955;230:1–19. a turbulent medium. Applied Scientific Research 1993;51:519–24.
[117] De Joannon M, Ragucci R, Cavaliere A. Laser excited emission and chem- [147] Lakshmisha KN, Rogg B, Bray KNC. PDF modelling of autoignition in non-
iluminescence from autoigniting spray. Combustion Science and Technology premixed turbulent flows. Combustion Science and Technology 1995;105:
2000;155:129–47. 229–43.
[118] Gordon RL. PhD thesis. University of Sydney; 2007. [148] Zhu M, Bray KNC, Rogg B. PDF modelling of spray autoignition in high
[119] Ayoola BO, Balachandran R, Frank JH, Mastorakos E, Kaminski CF. Spatially pressure turbulent flows. Combustion Science and Technology
resolved heat release rate measurements in turbulent premixed flames. 1996;120:357–79.
Combustion and Flame 2006;144:1–16. [149] Correa SM, Dean AJ. Turbulent broadening of autoignition limits. Proceedings
[120] Blouch JD, Sung CJ, Fotache CG, Law CK. Turbulent ignition of non-premixed of the Combustion Institute 1994;25:1293–9.
hydrogen by heated counterflowing atmospheric air. Proceedings of the [150] Masri AR, Cao R, Pope SB, Goldin GM. PDF calculations of turbulent lifted
Combustion Institute 1998;27:1221–8. flames of H2/N2 fuel issuing into a vitiated co-flow. Combustion Theory and
[121] Blouch JD, Law CK. Effects of turbulence on nonpremixed ignition of Modelling 2004;8:1–22.
hydrogen in heated counterflow. Combustion and Flame 2003;132:512–22. [151] Cao RR, Pope SB, Masri AR. Turbulent lifted flames in a vitiated coflow
[122] Lovas T, Lowe A, Cant RS, Mastorakos E. Three-dimensional direct numerical investigated using joint PDF calculations. Combustion and Flame
simulations of autoignition in turbulent non-premixed flows with simple 2005;142:438–53.
and complex chemistry. Proceedings of FEDSM2006 2006. ASME paper [152] Gkagkas K, Lindstedt RP. Transported PDF modelling with detailed chemistry
FEDSM2006-98109. of pre- and auto-ignition in CH4/air mixtures. Proceedings of the Combustion
[123] Markides CN, Mastorakos E. Measurements of the statistical distribution of Institute 2007;31:1559–66.
the scalar dissipation rate in turbulent axisymmetric plumes. Flow, Turbu- [153] Gordon RL, Masri AR, Pope SB, Goldin GM. Transport budgets in turbulent
lence and Combustion 2008;81:221–34. lifted flames of methane autoigniting in a vitiated co-flow. Combustion and
[124] Koss HJ, Brüggemann D, Wiartalla A, Bäcker H, Breuer A. Investigations of the Flame 2007;151:495–511.
influence of turbulence and type of fuel on the evaporation and mixture [154] Blouch JD, Chen J-Y, Law CK. A joint scalar PDF study of nonpremixed
formation in fuel sprays. Idea project. University of Aachen; 1992. hydrogen ignition. Combustion and Flame 2003;135:209–25.
[125] Mizutani Y, Nakabe K, Chung JD. Effects of turbulent mixing on spray ignition. [155] Sabel’nikov V, Gorokhovski M, Baricault N. The extended IEM mixing model
Proceedings of the Combustion Institute 1990;23:1455–60. in the framework of the composition PDF approach: application to diesel
[126] Kobori S, Kamimoto T, Kosaka H. Ignition, combustion and emissions in spray combustion. Combustion Theory and Modelling 2006;10:155–69.
a DI diesel engine equipped with a micro-hole nozzle. SAE Paper 1996; [156] Maigaard P, Mauss F, Kraft M. Homogeneous charge compression ignition
960321. engine: a simulation study on the effects of inhomogeneities. Journal of
[127] Markides CN, De Paola G, Mastorakos E. Measurements and simulations of Engineering for Gas Turbines and Power 2003;125:466–71.
mixing and autoignition of an n-heptane plume in a turbulent flow of heated [157] Zhang YZ, Kung EH, Haworth DC. A PDF method for multidimensional
air. Experimental Thermal and Fluid Science 2007;31:393–401. modeling of HCCI engine combustion: effects of turbulence/chemistry
[128] Naber JD, Siebers DL, Di Julio SS, Westbrook CK. Effects of natural gas interactions on ignition timing and emissions. Proceedings of the Combus-
composition on ignition delay under diesel conditions. Combustion and tion Institute 2005;30:2763–71.
Flame 1994;99:192–200. [158] Lee C-W, Mastorakos E. Numerical simulations of homogeneous charge
[129] Haessler H, Fast G, Kuhn D, Class AG, Bockhorn H. Auto-ignition of dime- compression ignition engines with high levels of residual gas. International
thylether at high pressure. In Work-In-Progress Poster Abstracts. 31st Inter- Journal of Engine Research 2007;8:63–78.
national Symposium on Combustion; 2006. [159] Jones WP, Navarro-Martinez S. Large eddy simulation of autoignition with
[130] Fast G, Kuhn D, Class AG, Maas U. Auto-ignition during instationary jet a subgrid probability density function method. Combustion and Flame
evolution of dimethyl ether (DME) is a high-pressure environment. 2007;150:170–87.
Combustion and Flame. Submitted for publication. [160] Jones WP, Navarro-Martinez S, Röhl O. Large eddy simulation of hydrogen
[131] Baritaud TA, Heinze TA, Le Coz JF. Spray and self-ignition visualization in a DI autoignition with a probability density function method. Proceedings of the
diesel engine. SAE Paper 1994;940681. Combustion Institute 2007;31:1765–71.
[132] Golub VV, Baklanov DI, Bazhenova TV, Bragin MV, Golovastov SV, Ivanov MF, [161] Jones WP, Navarro-Martinez S. Study of hydrogen auto-ignition in a turbulent
et al. Shock-induced ignition of hydrogen gas during accidental or technical air co-flow using a large eddy simulation approach. Computers and Fluids
opening of high-pressure tanks. Journal of Loss and Prevention in the Process 2008;37:802–8.
Industries 2007;20:439–46. [162] Barths H, Hasse C, Bikas G, Peters N. Simulation of combustion in direct
[133] Echekki T, Chen JH. High-temperature combustion in autoigniting non- injection diesel engines using a Eulerian particle flamelet model. Proceedings
homogeneous hydrogen/air mixtures. Proceedings of the Combustion Insti- of the Combustion Institute 2000;28:1161–8.
tute 2002;29:2061–8. [163] Pitsch H, Fedotov S. Investigation of scalar dissipation rate fluctuations in
[134] Seiser R, Frank JH, Liu S, Chen JH, Sigurdsson RJ, Seshadri K. Ignition of non-premixed turbulent combustion using a stochastic approach. Combus-
hydrogen in unsteady nonpremixed flows. Proceedings of the Combustion tion Theory and Modelling 2001;5:41–57.
Institute 2005;30:423–30. [164] Blanquart G, Pitsch H. Modeling autoignition in non-premixed turbulent
[135] Markides CN, Mastorakos E. Flame propagation following the autoignition of combustion using a stochastic flamelet approach. Proceedings of the
axisymmetric hydrogen, acetylene and normal-heptane plumes in turbulent Combustion Institute 2005;30:2745–53.
coflows of hot air. Journal of Engineering for Gas Turbines and Power [165] Hasse C, Peters N. A two mixture fraction flamelet model applied to split
2008;130(011502). injections in a DI diesel engine. Proceedings of the Combustion Institute
[136] Cheng TS, Wehrmeyer JA, Pitz RW. Conditional analysis of lifted hydrogen jet 2005;30:2755–62.
diffusion flame experimental data and comparison to laminar flame solu- [166] Bruel P, Rogg B, Bray KNC. On auto-ignition in laminar and turbulent non-
tions. Combustion and Flame 2007;150:340–54. premixed systems. Proceedings of the Combustion Institute 1990;23:759–66.
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 95

[167] Zhang Y, Rogg B, Bray KNC. 2-D simulation of turbulent autoignition with [198] Stauch R, Maas U. The ignition of methanol droplets in a laminar convective
transient laminar flamelet source term closure. Combustion Science and environment. Combustion and Flame 2008;153:45–57.
Technology 1995;105:211–27. [199] Khan QS, Baek SW, Ghassemi H. On the autoignition and combustion char-
[168] Chang C-S, Zhang Y, Bray KNC, Rogg B. Modelling and simulation of auto- acteristics of kerosene droplets at elevated pressure and temperature.
ignition under simulated diesel-engine conditions. Combustion Science and Combustion Science and Technology 2007;179:2437–51.
Technology 1996;113–114:205–19. [200] Wang C-H, Shy K-H, Lieu L-C. An experimental investigation on the
[169] Lehtiniemi H, Mauss F, Balthasar M, Magnusson I. Modeling diesel spray ignition delay of fuel droplets. Combustion Science and Technology
ignition using detailed chemistry with a progress variable approach. 1996;118:63–78.
Combustion Science and Technology 2006;178:1977–97. [201] Tanabe M, Kono M, Sato J, Koenig J, Eigenbrod G, Dinkelacher F, et al. Two
[170] Pires Da Cruz FA, Baritaud TA, Poinsot TJ. Self-ignition and combustion stage ignition of n-heptane isolated droplets. Combustion Science and
modeling of initially non-premixed turbulent systems. Combustion and Technology 1995;108:103–19.
Flame 2001;124:65–81. [202] Mullins BP. The spontaneous combustion of fuels injected into a hot gas
[171] Tap FA, Hilbert R, Thevenin D, Veynante D. A generalized flame surface stream. Symposium on Combustion, Flame and Explosion Phenomena
density modelling approach for the auto-ignition of a turbulent non-pre- 1949;3:704–13.
mixed system. Combustion Theory and Modelling 2004;8:165–93. [203] Penner SS, Mullins BP. Explosions, detonations, flammability and ignition,
[172] Tap FA, Veynante D. Simulation of flame lift-off on a diesel jet using AGARDograph 31. Pergamon Press; 1959.
a generalized flame surface density modeling approach. Proceedings of the [204] Bellan J, Harstad K. Ignition of nondilute clusters of drops in convective flows.
Combustion Institute 2005;30:919–26. Combustion Science and Technology 1987;53:75–87.
[173] Michel J-B, Colin O, Veynante D. Modeling ignition and chemical structure of [205] Segawa D, Yoshida M, Nakaya S, Kadota T. Autoignition and early flame
partially premixed turbulent flames using tabulated chemistry. Combustion behaviour of a spherical cluster of 49 monodispersed droplets. Proceedings
and Flame 2008;152:80–99. of the Combustion Institute 2007;31:2149–56.
[174] Domingo P, Vervisch L, Veynante D. Large-eddy simulation of a lifted [206] Wong S-C, Chang J-C, Yang J-C. Autoignition of droplets in nondilute
methane jet flame in a vitiated coflow. Combustion and Flame 2008;152: monodisperse clouds. Combustion and Flame 1993;94:397–406.
415–32. [207] Dwyer HA, Stapf P, Maly R. Unsteady vaporization and ignition of a three-
[175] Duwig C, Fuchs L. Large eddy simulation of a H2/N2 lifted flame in a vitiated dimensional droplet array. Combustion and Flame 2000;121:181–94.
co-flow. Combustion Science and Technology 2008;180:453–80. [208] Wang Y, Rutland CJ. Effects of temperature and equivalence ratio on the
[176] Klimenko AY, Bilger RW. Conditional moment closure for turbulent ignition of n-heptane fuel spray in turbulent flow. Proceedings of the
combustion. Progress in Energy and Combustion Science 1999;25:595–687. Combustion Institute 2005;30:893–900.
[177] Kim SH, Huh KY, Fraser RA. Modeling autoignition of a turbulent methane jet [209] Wang Y, Rutland CJ. DNS study of the ignition of n-heptane fuel spray under
by the conditional moment closure model. Proceedings of the Combustion high pressure and lean conditions. Journal of Physics: Conference Series
Institute 2000;28:185–91. 2005;16:124–8.
[178] Kim WT, Huh KY. Numerical simulation of spray autoignition by the first- [210] Wang Y, Rutland CJ. Direct numerical simulation of ignition in turbulent n-
order conditional moment closure model. Proceedings of the Combustion heptane liquid-fuel spray jets. Combustion and Flame 2007;149:353–65.
Institute 2002;29:569–76. [211] Schroll P, Wandel AP, Cant RS, Mastorakos E. Direct numerical simulations of
[179] De Paola G, Mastorakos E, Wright YM, Boulouchos K. Diesel engine simula- autoignition in turbulent two-phase flows. Proceedings of the Combustion
tions with multi-dimensional conditional moment closure. Combustion Institute, In press.
Science and Technology 2008;180:883–99. [212] Vogel S, Hasse C, Gronki J, Andersson S, Peters N, Wolfrum J, et al. Numerical
[180] Grout RW, Bushe WK, Blair C. Predicting the ignition delay of turbulent simulation and laser-based imaging of mixture formation, ignition, and soot
methane jets using conditional source-term estimation. Combustion Theory formation in a diesel spray. Proceedings of the Combustion Institute
and Modelling 2007;11:1009–28. 2005;30:2029–36.
[181] Mastorakos E, Bilger RW. Second-order conditional moment closure for the [213] Demoulin FX, Borghi R. Assumed PDF modeling of turbulent spray combus-
autoignition of turbulent flows. Physics of Fluids 1998;10:1246–8. tion. Combustion Science and Technology 2000;158:249–71.
[182] De Paola G, Kim IS, Mastorakos E. Second-order conditional moment [214] Demoulin FX, Borghi R. Modeling of turbulent spray combustion with
closure simulations of autoignition of an n-heptane plume in a turbulent application to diesel like environment. Combustion and Flame
coflow of heated air. Flow, Turbulence and Combustion. In press. 2002;129:281–93.
[183] Kim IS, Mastorakos E. Simulations of turbulent non-premixed counterflow [215] Ge H-W, Gutheil E. Simulation of a turbulent spray flame using coupled PDF
flames with first-order conditional moment closure. Flow, Turbulence and gas phase and spray flamelet modeling. Combustion and Flame
Combustion 2006;76:133–62. 2008;153:173–85.
[184] Richardson ES, Chakraborty N, Mastorakos E. Analysis of direct numerical [216] Colin O, Benkenida A. A new scalar fluctuation model to predict mixing in
simulations of ignition fronts in turbulent non-premixed flames in the evaporating two-phase flows. Combustion and Flame 2003;134:207–27.
context of conditional moment closure. Proceedings of the Combustion [217] Desjardins O, Pitsch H. Modeling effect of spray evaporation on turbulent
Institute 2007;31:1683–90. combustion. Number paper ICLASS06-084, Kyoto, Japan; August 27–
[185] De Paola G. Personal communication. 2007. September 1 2006. ICLASS-2006.
[186] Patwardhan SS, Santanu De, Lakshmisha KN, Raghunandan BN. CMC simu- [218] Ronney PD. Laser versus conventional ignition of flames. Optical Engineering
lations of lifted turbulent jet flame in a vitiated coflow. Proceedings of the 1994;33:510–21.
Combustion Institute. In press. [219] Bradley D, Sheppard CGW, Suardjaja IM, Woolley R. Fundamentals of high-
[187] Navarro-Martinez S, Kronenburg A. LES-CMC simulations of a lifted methane energy spark ignition with lasers. Combustion and Flame 2004;138:55–77.
flame. Proceedings of the Combustion Institute. In press. [220] Beduneau J-L, Kim B, Zimmer L, Ikeda Y. Measurements of minimum ignition
[188] Echekki T. Stochastic modeling of autoignition in turbulent non-homoge- energy in premixed laminar methane/air flow by using laser induced spark.
neous hydrogen-air mixtures. International Journal of Hydrogen Energy Combustion and Flame 2003;132:653–65.
2008;33:2596–603. [221] Murase E, Ono S, Hanada K, Oppenheim AK. Initiation of combustion in lean
[189] Danby SJ, Echekki T. Proper orthogonal decomposition analysis of auto- mixtures by flame jets. Combustion Science and Technology 1996;113–
ignition simulation data of nonhomogeneous hydrogen-air mixtures. 114:167–77.
Combustion and Flame 2006;144:126–38. [222] Pitt PL, Ridley JD, Clements RM. An ignition system for ultra lean mixtures.
[190] Gourara A, Roger F, Wang HY, Most J-M. Assessment of ignition hazard in Combustion Science and Technology 1984;35:277–85.
turbulent flammable gas mixers combining a Lagrangian approach and large [223] Sadanandan R, Markus D, Schiessl R, Maas U, Olofsson J, Seyfried H, et al.
eddy simulation. Combustion and Flame 2006;144:592–604. Detailed investigation of ignition by hot gas jets. Proceedings of the
[191] Law CK, Chung SH. An ignition criterion for droplets in sprays. Combustion Combustion Institute 2007;31:719–26.
Science and Technology 1980;22:17–26. [224] Cote T, Ridley JD, Clements RM, Smy PR. The ignition characteristics of
[192] Cuoci A, Mehl M, Buzzi-Ferraris G, Faravelli T, Manca D, Ranzi E. Autoignition igniters at sub-atmospheric pressures. Combustion Science and Technology
and burning rates of fuel droplets under microgravity. Combustion and Flame 1986;48:151–62.
2005;143:211–26. [225] Campbell CS, Egolfopoulos FN. Kinetics paths to radical-induced ignition of
[193] Gutheil E. Numerical analysis of the autoignition of methanol, ethanol, n- methane/air mixtures. Combustion Science and Technology 2005;177:
heptane and n-octane sprays with detailed chemistry. Combustion Science 2275–98.
and Technology 1995;105:265–78. [226] Marble FE, Adamson Jr TC. Ignition and combustion in a laminar mixing
[194] Yang J-R, Wong S-C. On the suppression of negative temperature coefficient zone. Jet Propulsion 1954;March–April:85–94.
(NTC) in autoignition of n-heptane droplets. Combustion and Flame [227] Kono M, Hatori K, Iinuma K. Investigation on ignition ability of composite
2003;132:475–91. sparks in flowing mixtures. Proceedings of the Combustion Institute
[195] Moriue O, Mikami M, Kojima N, Eigenbrod C. Numerical simulations of the 1984;20:133–40.
ignition of n-heptane droplets in the transition diameter range from [228] Ishii K, Tsukamoto T, Ujiie Y, Kono M. Analysis of ignition mechanism of
heterogeneous to homogeneous ignition. Proceedings of the Combustion combustible mixtures by composite sparks. Combustion and Flame
Institute 2005;30:1973–80. 1992;91:153–64.
[196] Stauch R, Lipp S, Maas U. Detailed numerical simulations of the autoignition [229] Kravchik T, Sher E, Heywood JB. From spark ignition to flame initiation.
of single n-heptane droplets in air. Combustion and Flame 2006;145:533–42. Combustion Science and Technology 1995;108:1–30.
[197] Stauch R, Maas U. The auto-ignition of single n-heptane/iso-octane droplets. [230] Thiele M, Warnatz J, Maas U. Geometrical study of spark ignition in two
International Journal of Heat and Mass Transfer 2007;50:3047–53. dimensions. Combustion Theory and Modelling 2000;4:413–34.
96 E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97

[231] Thiele M, Warnatz J, Dreizler A, Lindenmaier S, Schiessl R, Maas U, et al. Spark [262] Pasquier N, Lecordier B, Trinité M, Cessou A. An experimental investigation of
ignited hydrogen/air mixtures: two dimensional detailed modeling and laser flame propagation through a turbulent stratified mixture. Proceedings of the
based diagnostics. Combustion and Flame 2002;128:74–87. Combustion Institute 2007;31:1567–74.
[232] Strehlow RA. Fundamentals of combustion. International Textbook [263] Drake MC, Fansler TD, Lippert AM. Stratified-charge combustion: modeling
Company; 1968. and imaging of a spray-guided direct-injection spark-ignition engine.
[233] Swett Jr. CC, Donlon RH. Spark ignition of flowing gases III – effect of Proceedings of the Combustion Institute 2005;30:2683–91.
turbulence promoter on energy required to ignite a propane-air mixture. [264] Jiménez C, Cuenot B, Poinsot T, Haworth D. Numerical simulations and
NACA Research Memorandum; 1953. p. RM E52J28. modeling for lean stratified propane-air flames. Combustion and Flame
[234] Bradley D, Lung FK-K. Spark ignition and the early stages of turbulent flame 2002;128:1–21.
propagation. Combustion and Flame 1987;69:71–93. [265] Phillips H. Flame in a buoyant methane flame. Proceedings of the Combus-
[235] Ballal DR, Lefebvre AH. Ignition and flame quenching in flowing gaseous tion Institute 1965;10:1277–83.
mixtures. Proceedings of the Royal Society of London A 1977;357:163–81. [266] Fernandez-Tarrazo E, Vera M, Liñán A. Liftoff and blowoff of a diffusion flame
[236] Huang CC, Shy SS, Liu CC, Yan YY. A transition on minimum ignition energy between parallel streams of fuel and air. Combustion and Flame
for lean turbulent methane combustion in flamelet and distributed regimes. 2006;144:261–76.
Proceedings of the Combustion Institute 2007;31:1401–9. [267] Briones AM, Aggarwal SK, Katta VR. Effects of H2 enrichment on the prop-
[237] Ballal DR, Lefebvre AH. Flame quenching in turbulent flowing gaseous agation characteristics of CH4-air triple flames. Combustion and Flame
mixtures. Proceedings of the Combustion Institute 1977;16:1689–98. 2008;153:367–83.
[238] Ballal DR, Lefebvre AH. A general model of spark ignition for gaseous and [268] Cha MS, Ronney PD. Propagation rates of nonpremixed edge flames.
liquid fuel-air mixtures. Proceedings of the Combustion Institute Combustion and Flame 2006;146:312–28.
1981;18:1737–46. [269] Buckmaster J. Egde-flames and their stability. Combustion Science and
[239] Klein M, Chakraborty N, Cant RS. Effects of turbulence on self-sustained Technology 1996;115:41–68.
combustion in premixed flame kernels: a direct numerical simulation study. [270] Chakraborty N, Mastorakos E. Numerical investigation of edge flame prop-
Flow, Turbulence and Combustion, In press. agation characteristics in turbulent mixing layers. Physics of Fluids
[240] Hamai K, Kawajiri H, Ishizuka T, Nakai M. Combustion fluctuation mech- 2006;18:105103.
anism involving cycle-to-cycle spark ignition variation due to gas flow [271] Chakraborty N, Mastorakos E, Cant RS. Effects of turbulence on spark ignition
motion in S.I. engines. Proceedings of the Combustion Institute 1986; in inhomogeneous mixtures: a direct numerical simulation (DNS) study.
21:505–12. Combustion Science and Technology 2007;179:293–317.
[241] Aleiferis PG, Hardalupas Y, Taylor AMKP, Ishii K, Urata Y. Flame chem- [272] Chakraborty N, Mastorakos E. Direct numerical simulations of localised
iluminescence studies of cyclic combustion variations and air-to-fuel ratio of forced ignition in turbulent mixing layers: the effects of mixture fraction and
the reacting mixture in a lean-burn stratified-charge spark-ignition engine. its gradient. Flow, Turbulence and Combustion 2008;80:155–86.
Combustion and Flame 2004;136:72–90. [273] Poinsot T, Candel S, Trouve A. Applications of direct numerical simulation to
[242] Pischinger S, Heywood JB. How heat losses to the spark plug electrodes affect premixed turbulent combustion. Progress in Energy and Combustion Science
flame kernel development in an SI-engine. SAE Technical Paper 1995;21:531–76.
1990;892083:2006–14. [274] Barbosa S, Scouflaire P, Ducruix S, Gaborel G. Comparisons between spark
[243] Rashkovsky SA. Spark ignition of ill-mixed gases. In: First Mediterranean plug and laser ignition in a gas turbine combustor. In: Third European
Combustion Symposium, Antalya, Turkey; 1999. Combustion Metting, Crete, Greece; 2007.
[244] Richardson ES, Mastorakos E. Numerical investigation of forced ignition in [275] Gurliat O, Schmidt V, Haidn OJ, Oschwald M. Ignition of cryogenic H2/LOX
laminar counterflow non-premixed methane-air flames. Combustion Science sprays. Aerospace Science and Technology 2003;7:517–31.
and Technology 2007;179:21–37. [276] McManus K, Aguerre F, Yip B, Candel S. Analysis of the ignition sequence of
[245] Phuoc TX, White CM, McNeill DH. Laser spark ignition of a jet diffusion flame. a multiple injector combustor using PLIF imaging. In: Third International
Optics and Lasers in Engineering 2002;38:217–32. Symposium on Special Topics in Chemical Propulsion: Non-intrusive
[246] Phuoc X, Chen RH. Use of laser-induced spark for studying ignition stability Combustion Diagnostics, Scheveningen; 1993.
and unburned hydrogen escaping from laminar diluted hydrogen diffusion [277] Nicholson HM, Field JP. Some experimental techniques for the investigation
jet flames. Optics and Lasers in Engineering 2007;45:834–42. of the mechanism of flame stabilization in the wakes of bluff bodies.
[247] Qin X, Choi CW, Mukhopadhyay A, Puri IK, Aggarwal SK, Katta VR. Triple Symposium on Combustion, Flame and Explosion Phenomena 1949;3:44–68.
flame propagation and stabilization in a laminar axisymmetric jet. [278] Naegeli DW, Dodge LG. Ignition study in a gas turbine combustor. Combus-
Combustion Theory and Modelling 2004;8:293–314. tion Science and Technology 1991;80:165–84.
[248] Birch AD, Brown DR, Dodson MG. Ignition probabilities in turbulent mixing [279] Read RW, Rogerson JW, Hochgreb S. Relight imaging at low temperature, low
flows. Proceedings of the Combustion Institute 1981;18:1775–80. pressure conditions. AIAA Paper 2008. IAA-2008-0957.
[249] Smith MTE, Birch AD, Brown DR, Fairweather M. Studies of ignition and [280] Alvani RF, Fairweather M. Ignition characteristics of turbulent jet flows.
flame propagation in turbulent jets of natural gas, propane and a gas with Chemical Engineering Research and Design 2002;80:917–23.
a high hydrogen content. Proceedings of the Combustion Institute [281] Alvani RF, Fairweather M. Modelling of the ignition characteristics of flam-
1986;21:1403–8. mable jets using a k–3–g turbulence model. Turbulence, Heat and Mass
[250] Birch AD, Brown DR, Cook DK, Hargrave GK. Flame stability in under- Transfer 4, New York; 2003. p. 911–18.
expanded natural-gas jets. Combustion Science and Technology [282] Veynante D, Lacas F, Candel SM. Numerical simulation of the transient
1988;58:267–80. ignition regime of a turbulent diffusion flame. AIAA Journal 1991;29:848–51.
[251] Ahmed SF, Mastorakos E. Spark ignition of lifted turbulent jet flames. [283] Colin O, Benkenida A, Angelberger C. 3D modeling of mixing, ignition and
Combustion and Flame 2006;146:215–31. combustion phenomena in highly stratified gasoline engines. Oil and Gas
[252] McCraw JL, Moore NJ, Lyons KM. Observations on upstream flame propaga- Science and Technology – Revue de l’Institut Francais du Petrole
tion in the ignition of hydrocarbon jets. Flow, Turbulence and Combustion 2003;58:47–62.
2007;79:83–97. [284] Helie J, Trouve A. A modified coherent flame model to describe turbulent
[253] Swain MR, Filoso PA, Swain MN. An experimental investigation into the flame propagation in mixtures with variable composition. Proceedings of the
ignition of leaking hydrogen. International Journal of Hydrogen Energy Combustion Institute 2000;28:193–201.
2007;32:287–95. [285] Wallesten J, Lipatnikov A, Chomiak J. Modeling of stratified combustion in
[254] Ahmed SF, Balachandran R, Mastorakos E. Measurements of ignition proba- a direct-injection spark-ignition engine accounting for complex chemistry.
bility in turbulent non-premixed counterflow flames. Proceedings of the Proceedings of the Combustion Institute 2002;29:703–9.
Combustion Institute 2007;31:1507–13. [286] Ma CY, Mahmud T, Fairweather M, Hampartsoumian E, Gaskell PH. Prediction
[255] El-Rabii H, Zahringer K, Rolon J-C, Lacas F. Laser ignition in a lean premixed of lifted, non-premixed turbulent flames using a mixedness-reactedness
prevaporized injector. Combustion Science and Technology 2004;176: flamelet model with radiation heat loss. Combustion and Flame
1391–417. 2002;128:60–73.
[256] Zimmer L, Okai K, Kurosawa Y. Laser ignition and plasma spectroscopy in [287] Bray K, Domingo P, Vervisch L. Role of the progress variable in models for
non-premixed hydrogen-air burner. In: ICDERS 21, Poitiers, France; July 2007. partially premixed turbulent combustion. Combustion and Flame
[257] Zimmer L, Tachibana S. Laser induced plasma spectroscopy for local equiv- 2005;141:431–7.
alence ratio measurements in an oscillating combustion environment. [288] Boileau M. Simulation aux grandes échelles de l’allumage diphasique des
Proceedings of the Combustion Institute 2007;31:737–45. foyers aéronautiques – TH/CFD/07/103. PhD thesis. Institut Polytechnique
[258] Müller CM, Breitbach H, Peters N. Partially premixed turbulent flame prop- National de Toulouse; 2007.
agation in jet flames. Proceedings of the Combustion Institute [289] Boileau M, Staffelbach G, Cuenot B, Poinsot T, Berat C. LES of an ignition sequence
1994;25:1099–106. in a full helicopter combustor. Combustion and Flame 2008;154:2–22.
[259] Chen M, Herrmann M, Peters N. Flamelet modeling of lifted turbulent [290] Peters JE, Mellor AM. An ignition model for quiescent fuel sprays. Combus-
methane/air and propane/air jet diffusion flames. Proceedings of the tion and Flame 1980;38:65–74.
Combustion Institute 2000;28:167–74. [291] Danis AM, Namer I, Cernansky NP. Droplet size and equivalence ratio effects
[260] Renou B, Samson E, Boukhalfa A. An experimental study of freely propa- on spark ignition of monodisperse n-heptane and methanol sprays.
gating turbulent propane/air flames in stratified inhomogeneous mixtures. Combustion and Flame 1988;74:285–94.
Combustion Science and Technology 2004;176:1867–90. [292] Dietrich DL, Cernansky NP, Somashekara MB, Namer I. Spark ignition of
[261] Kang T, Kyritsis DC. Methane flame propagation in compositionally stratified a bidisperse, n-decane fuel spray. Proceedings of the Combustion Institute
gases. Combustion Science and Technology 2005;177:2191–210. 1990;23:1383–9.
E. Mastorakos / Progress in Energy and Combustion Science 35 (2009) 57–97 97

[293] Singh AK, Polymeropoulos CE. Spark ignition of aerosols. Proceedings of the [302] Wandel AP, Chakraborty N, Mastorakos E. Direct numerical simulation of
Combustion Institute 1986;21:513–9. turbulent flame expansion in fine sprays. Proceedings of the Combustion
[294] Ballal DR, Lefebvre AH. Ignition of liquid fuel sprays at subatmospheric Institute. In press.
pressures. Combustion and Flame 1978;31:115–26. [303] Reveillon J, Demoulin FX. Evaporating droplets in turbulent reacting flows.
[295] Lee K-P, Wang S-H, Wong S-C. Spark ignition characteristics of monodisperse Proceedings of the Combustion Institute 2007;31:2319–26.
multicomponent fuel sprays. Combustion Science and Technology 1996;113– [304] Myers GD, Lefebvre AH. Flame propagation in heterogeneous mixtures of
114:493–502. fuel drops and air. Combustion and Flame 1986;66:193–210.
[296] Lawes M, Lee Y, Mokhtar AS, Woolley R. Laser ignition of iso-octane air [305] Richards GA, Lefebvre AH. Turbulent flame speeds of hydrocarbon fuel
aerosols. Combustion Science and Technology 2008;180:296–313. droplets in air. Combustion and Flame 1989;78:299–307.
[297] Ballal DR, Lefebvre AH. Flame propagation in heterogeneous mixtures of fuel [306] Reveillon J, Vervisch L. Analysis of weakly turbulent dilute-spray flames
droplets, fuel vapor and air. Proceedings of the Combustion Institute and spray combustion regimes. Journal of Fluid Mechanics 2005;537:
1981;18:321–8. 317–47.
[298] Zhu M, Rogg B. Modelling and simulation of sprays in laminar flames. [307] Cheriyan GK, Krallis K, Weinberg FJ. Adapting continuous-flow plasma jets
Meccanica 1996;31:177–93. for intermittent ignition in gas turbine combustors. Combustion Science and
[299] Niioka T. Flame propagation in sprays and particle clouds of less volatile Technology 1990;70:171–85.
fuels. Combustion Science and Technology 2005;177:1167–82. [308] Ouarti N. Modelisation de l’allumage d’un brouillard de carburant dans un
[300] Greenberg JB, Kalma A. A study of stretch in premixed spray flames. foyer de turbomachine. PhD thesis. Ecole Nationale Superieur de l’Aer-
Combustion and Flame 2000;123:421–9. onautique et de l’Espace, ONERA; 2004.
[301] Greenberg JB. Finite-rate evaporation and droplet drag effects in spherical [309] Quintilla V, Lecourt R, Lavergne G. Developpement d’un modele en vue de la
flame front propagation through a liquid fuel mist. Combustion and Flame prediction du rallumage d’un foyer de turboreacteur a haute altitude.
2007;148:187–97. Comptes Rendus Mecanique 2002;330:811–8.

You might also like