Vous êtes sur la page 1sur 8

Available online at www.sciencedirect.

com
Available online at www.sciencedirect.com
ScienceDirect
ScienceDirect
Energy
Available
Available Procedia
online
online 00 (2017) 000–000
atatwww.sciencedirect.com
www.sciencedirect.com
Energy Procedia 00 (2017) 000–000 www.elsevier.com/locate/procedia
ScienceDirect
ScienceDirect
www.elsevier.com/locate/procedia

Energy
EnergyProcedia 126
Procedia 00(201709) 826–833
(2017) 000–000
www.elsevier.com/locate/procedia
72nd Conference of the Italian Thermal Machines Engineering Association, ATI2017, 6-8
72nd Conference of the ItalianSeptember
Thermal Machines Engineering
2017, Lecce, Italy Association, ATI2017, 6-8
September 2017, Lecce, Italy
Modelling soot production and thermal radiation for turbulent
Modelling
The soot production
15th International and thermal
Symposium radiation
on District Heating andfor turbulent
Cooling
diffusion flames
diffusion flames
AssessingLorenzo
the feasibility of using the heat demand-outdoor
Mazzei*, Stefano Puggelli, Davide Bertini,
temperature Lorenzo
function
Danielefor
Mazzei*,a long-term
Stefano
Pampaloni district
Puggelli,
and Antonio Davideheat demand forecast
Bertini,
Andreini
Daniele Pampaloni and Antonio Andreini
Department
a,b,c of Industrial Engineering, University of Florence, Via
b S. Marta 3, Firenze – 50139, Italy
I. Andrić *, A. Pina , P. Ferrão , J. Fournier ., B. Lacarrière , O. Le Correc
a a c
Department of Industrial Engineering, University of Florence, Via S. Marta 3, Firenze – 50139, Italy
a
IN+ Center for Innovation, Technology and Policy Research - Instituto Superior Técnico, Av. Rovisco Pais 1, 1049-001 Lisbon, Portugal
Abstract b
Veolia Recherche & Innovation, 291 Avenue Dreyfous Daniel, 78520 Limay, France
Abstract c
Département Systèmes Énergétiques et Environnement - IMT Atlantique, 4 rue Alfred Kastler, 44300 Nantes, France
This paper presents a systematic investigation on the strategies required for modelling soot when performing CFD
This paper presents
simulations a systematic
of turbulent flames. Theinvestigation on the
first test case strategies
consists required
in a 3D for modelling
enclosure containingsoot when performing
a mixture of N 2, CO2, CFD
H2O
simulations
and of turbulent
soot (Coelho, 2003).flames.
ResultsThe first test
obtained withcase
the consists in a 3D enclosure
gray implementation of thecontaining a mixture
Discrete Ordinate of N 2,inCO
Method 2, H2O
ANSYS
and sootare(Coelho,
Abstract
Fluent compared 2003). Results
against obtained
literature with
data, the graydifferent
assessing implementation of thefor
formulations Discrete Ordinate coefficient
the absorption Method in ANSYS
of soot.
Fluent are compared are
These considerations against literature
exploited data,reactive
through assessing
RANS different formulations
simulations for the
accounting absorption
for radiative coefficient
heat of soot.
transfer, with the
District
These
purpose heating
to networks
considerations
test different are commonly
aremodels
exploited sootaddressed
for through inprocess
reactive
formation the literature
RANS as one accounting
simulations
on a turbulent ofnon-premixed
the mostfor
effective solutions
radiative for decreasing
heat transfer,
kerosene-air flame the
with the
(Young et
greenhouse
purpose
al., gasdifferent
to test
1994). emissionsmodels
from thefor
building sector. These
soot formation systems
process on arequire highnon-premixed
turbulent investments which are returned
kerosene-air through
flame the heat
(Young et
sales.
al., Due to the changed climate conditions and building renovation policies, heat demand in the future could decrease,
1994).
©prolonging the investment
2017 The Authors. return
Published byperiod.
Elsevier Ltd.
©The2017 Thescope
main Authors. Published
paper isby
ofresponsibility
this Elsevier
to Elsevier
assess Ltd.
the feasibility of using
© 2017 The
Peer-review Authors.
under Published by
of the Ltd.
scientific committee of thethe heat
72nd demand –ofoutdoor
Conference temperature
the Italian Thermal function
MachinesforEngineering
heat demand
Peer-review
forecast. under
The responsibility
district of of thelocated
Alvalade, scientific committee
in Lisbon of the 72nd
(Portugal), nd Conference
used as aof the Italian
study. Thermal Machines Engineering
Peer-review
Association. under responsibility of the scientific committee of the 72was Conference ofcase
the Italian The district
Thermal is consisted
Machines of 665
Engineering
Association
buildings that vary in both construction period and typology. Three weather scenarios (low, medium, high) and three district
Association.
renovation scenarios were developed (shallow,
Keywords: Combustion; Soot; Radiative heat transfer; CFD. intermediate, deep). To estimate the error, obtained heat demand values were
comparedCombustion;
Keywords: with resultsSoot;
from a dynamic
Radiative heat heat demand
transfer; CFD. model, previously developed and validated by the authors.
The results showed that when only weather change is considered, the margin of error could be acceptable for some applications
1.(the error in annual demand was lower than 20% for all weather scenarios considered). However, after introducing renovation
Introduction
scenarios, the error value increased up to 59.5% (depending on the weather and renovation scenarios combination considered).
1. Introduction
The value of slope coefficient increased on average within the range of 3.8% up to 8% per decade, that corresponds to the
The emission of particulate matter, or soot, can be considered a side effect associated to combustion processes
decrease in the number of heating hours of 22-139h during the heating season (depending on the combination of weather and
The emission
characterized of particulate
by high temperature matter,
and rich or conditions.
soot, can beThe considered
production a side effect
of these associated
carbon to combustion
nanoparticles entails a processes
reduction
renovation scenarios considered). On the other hand, function intercept increased for 7.8-12.7% per decade (depending on the
characterized
incoupled
the efficiencyby high
of temperature
the combustion anddevices,
rich conditions.
as well The
as aproduction
detrimental ofeffect
these carbon
on nanoparticles
climate change entails
and a reduction
human health.
scenarios). The values suggested could be used to modify the function parameters for the scenarios considered, and
inimprove
the efficiency of the combustion devices,
the accuracy of heat demand estimations. as well as a detrimental effect on climate change and human health.

© 2017 The Authors. Published by Elsevier Ltd.


Peer-review under responsibility of the Scientific Committee of The 15th International Symposium on District Heating and
* Corresponding author. Tel.: +39-055-275-8712.
Cooling.
E-mail address: lorenzo.mazzei@htc.de.unifi.it
* Corresponding author. Tel.: +39-055-275-8712.
E-mail address: lorenzo.mazzei@htc.de.unifi.it
Keywords: Heat demand; Forecast; Climate change
1876-6102 © 2017 The Authors. Published by Elsevier Ltd.
Peer-review underThe
1876-6102 © 2017 responsibility of theby
Authors. Published scientific
Elsevier committee
Ltd. of the 72 nd Conference of the Italian Thermal Machines Engineering
Association.
Peer-review under responsibility of the scientific committee of the 72 nd Conference of the Italian Thermal Machines Engineering
Association.
1876-6102 © 2017 The Authors. Published by Elsevier Ltd.
Peer-review under responsibility of the Scientific Committee of The 15th International Symposium on District Heating and Cooling.
1876-6102 © 2017 The Authors. Published by Elsevier Ltd.
Peer-review under responsibility of the scientific committee of the 72nd Conference of the Italian Thermal Machines Engineering Association
10.1016/j.egypro.2017.08.266
Lorenzo Mazzei et al. / Energy Procedia 126 (201709) 826–833 827
2 Author name / Energy Procedia 00 (2017) 000–000

Therefore, a drastic reduction in emissions is mandatory, demanding for a deeper understanding in the formation
process, which is still an open issue in combustion chemistry research.
In the past years strategies for the reduction in soot emissions relied mainly on correlations, experience and trial-
and-error attempts. The increasingly diffusion of CFD is allowing to analyze and optimize combustion devices with
sufficient confidence. However, the uncertainty increases significantly when soot is concerned: the complexity of the
gas-phase chemistry and the numerous mechanisms involved in soot formation process, which are strongly coupled
with mixing and radiative heat transfer and depend on soot volume fraction itself, make the prediction of soot
emissions still a challenging task, leading to large errors in exhaust concentration even with small mispredictions in
the formation rates [1].
Different mechanisms are in fact involved in the determination of the final concentration at the exit of the
combustion device, such as inception, coagulation, surface growth and oxidation. Even though reasonable predictions
can be achieved on laminar diffusion flames [2], the scenario is further complicated when turbulent flames burning
common fuels such as diesel or kerosene are concerned, as the applicability of models developed for simple aliphatic
hydrocarbon fuels (such as methane) is questioned. Soot production in methane-air flames usually takes place from
small species (e.g. acetylene) that grow to form polyaromatic hydrocarbons (PAH) and eventually soot. However,
practical fuels such as kerosene contains aromatics, which significantly accelerate the soot formation process [3].
This paper presents a numerical investigation aimed at assessing different modelling strategies for soot formation
in a turbulent diffusive flame burning prevaporized kerosene. CFD RANS simulations were performed accounting for
combustion and radiation, whereas different soot models were compared.

2. Numerical modelling

2.1. Radiation modelling

The simulations presented in the present work are carried out in ANSYS Fluent, using the Discrete Ordinate
Method (DOM) to solve the Radiative Transfer Equation (RTE). The implementation of the DOM treats the medium
as gray, with a total emissivity coefficient of the gas εg calculated with a weighted sum of gray gases approach as:
𝑁𝑁𝑔𝑔
𝜀𝜀𝑔𝑔 = ∑𝑖𝑖=0 𝑎𝑎𝑖𝑖 (𝑇𝑇)[1 − 𝑒𝑒 −𝑘𝑘𝑖𝑖 𝑝𝑝𝑝𝑝 ] (1)

where ai and ki are the emissivity weighting factor and the absorption coefficient for the i-th gray gas component, p is
the sum of the partial pressures of all absorbing gases and s is the path length. In the range 0.001≤ps≤10.0 atm-m and
600≤T≤2400 K the code uses the coefficient values suggested by Smith et al. [4] for a mixture of CO2 and H2O
modelled with Ng=3 gray gases. Then, the absorption coefficient of the gas is calculated from the total emissivity gas
emissivity through the exploitation of Beer’s law as:
𝑙𝑙𝑙𝑙(1−𝜀𝜀𝑔𝑔 )
𝑎𝑎𝑔𝑔 = − (2)
𝑠𝑠

Absorption thus depends on the mean beam length of the path, which can be specified by the user or estimated as
s=3.6 V/A, where V is the fluid volume and A the total surface area of the fluid boundaries (wall, inlets and outlet,
excluding symmetry and periodicity).
The presence of soot is accounted adding a user defined scalar. Its radiative contribution is included as ag+s=ag+as,
where as is the absorption coefficient of soot. Different approaches were proposed in literature to estimate this value.
Given the small size of the primary soot particles, ranging from about 5 nm to 100 nm, the particles fall into the
Rayleigh regime, meaning that scattering can be assumed negligible compared to absorption and the spectral
absorption coefficient is given by kλ=C0fv/λ, where C0 is a constant depending only on the soot index of refraction, fv
is the volume fraction and λ the wavelength. The Planck-mean and the Rosseland mean absorption coefficients, valid
for the limiting conditions of optically thin and optically thick media. Their expressions differ only in a constant factor
by only 6%, therefore an average value was proposed by Felske and Tien [5] and reported by Modest [6]:

𝑘𝑘𝑚𝑚 = 3.72𝑓𝑓𝑣𝑣 𝐶𝐶0 𝑇𝑇/𝐶𝐶2 (3)


828 Lorenzo Mazzei et al. / Energy Procedia 126 (201709) 826–833
Author name / Energy Procedia 00 (2017) 000–000 3

where C2=1.4388 cm-K. The formula is valid only for very small soot particles, and the absorption coefficient will
increase for aggregates or if primary particle sizes exceed Rayleigh scattering limits. It is also worth underlining that
great uncertainty is ascribable to the value of the C0 constant, which is function of temperature, composition and
particle aggregation, in addition to wavelength. This explains the significant scatter in the values reported in literature,
ranging between 2 and 10, as a consequence of differences in the fuel employed and the typology of flame investigated,
with laboratory conditions that however differ significantly compared to the realistic applications.

2.2. Soot modelling

Soot models can be classified on the basis of their complexity and computational effort involved. According to
Kennedy [7], a distinction can be made in three categories: empirical, semi-empirical and detailed models. One of the
most known empirical model was proposed by Khan et al. [8] to predict soot emissions in diesel engines and,
neglecting particle growth and oxidation rates, estimates soot formation as a function of pressure, temperature and
equivalence ratio of unburnt gases. A more accurate description of the physics can be obtained with semi-empirical
models that include the chemical processes involved in soot formation, for example, expressing particle inception and
surface growth on the basis of PAH concentration. Lindstedt et al. [9] exploited simplified reactions to describe the
formation and growth of soot particles in laminar non-premixed ethylene and propane flames, using benzene and
acetylene as precursors. Other example are given by Fairweather et al. [10] and Brookes and Moss [11], which differ
for the choice of the precursor to soot nucleation, the model constants or the oxidation mechanism. High-fidelity
predictions in a wider range of fuels and operating conditions require to account for the dependency of soot inception
and surface growth on the PAH concentration. Detailed chemistry mechanisms are required to estimate PAH
formation, whereas additional pathways should be considered for surface reactions, as well as the agglomeration of
soot particles in complex agglomerate structures. This is accomplished with Monte Carlo models, sectional methods
or methods of moments (refer to Mueller et al. [12] for a description of these techniques). After this overview, the
soot models investigated in the present work are presented here below.

2.2.1. Moss-Brookes-Hall model


The Moss-Brookes-Hall model represents an extension of the formulation of the Brookes and Moss model [53],
which was developed and validated for methane flames. The original model consists in solving for two additional
transport equations for soot mass fraction Ysoot and normalized radical nuclei concentration b*nuc:
𝜇𝜇
𝜕𝜕
𝜕𝜕𝜕𝜕
(𝜌𝜌𝑌𝑌𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠 )+𝛻𝛻∙(𝜌𝜌𝑣𝑣⃗ 𝑌𝑌𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠 )=𝛻𝛻∙( 𝑡𝑡 𝛻𝛻𝑌𝑌𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠 )+
𝜎𝜎𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠
𝑑𝑑𝑑𝑑
𝑑𝑑𝑑𝑑
(4)
𝜕𝜕 ∗ )+𝛻𝛻∙(𝜌𝜌𝑣𝑣
(𝜌𝜌𝑏𝑏𝑛𝑛𝑛𝑛𝑛𝑛 ∗ )=𝛻𝛻∙( 𝜇𝜇𝑡𝑡 𝛻𝛻𝑏𝑏 ∗ )+𝑑𝑑𝑑𝑑
⃗ 𝑏𝑏𝑛𝑛𝑛𝑛𝑛𝑛  (5)
𝜕𝜕𝜕𝜕 𝜎𝜎𝑛𝑛𝑛𝑛𝑛𝑛 𝑛𝑛𝑛𝑛𝑛𝑛 𝑑𝑑𝑑𝑑

where M is soot mass concentration and N is soot particle number density. It is also worth specifying that the source
term for soot mass concentration dM/dt accounts for the different mechanisms of nucleation (source), surface growth
(source) and oxidation (sink), whereas the instantaneous production rate of soot particles dN/dt considers the
nucleation from the gas phase (source) and the coagulation in the free molecular regime (sink).
The Moss-Brooks-Hall model was reported by Wen et al. [13] and implements in the soot oxidation term the
dependence on O2, relevant at fuel lean conditions, besides the major contribution due to hydroxyl radical (OH), as
measured and modeled by Lee et al. [14]. Moreover, it is based on the model extension proposed by Hall et al. [15] to
include the contribution of two-ringed and three ringed aromatics (C10H7 and C14H10), benzene (C6H6) and the phenyl
radical (C6H5) in addition to acetylene (C2H2). Further modifications include the coagulation term and the surface
growth term, which are formulated similar to those used by Brookes and Moss model [11], but slightly modified
according to the model developed by Lindstedt [9]. Given the highly non-linear relationships among temperature,
species concentration and soot formation rate, a PDF is exploited to account for the interaction between turbulence
and chemistry. A more detailed description of the model can be found in ANSYS Fluent Theory Guide [16].

2.2.2. Method of Moments


To provide a significant improvement in the accuracy of the prediction of soot formation, detailed kinetic modelling
and resolution of particle-size distribution (PSD) are required. The PSD is used by the particle population balance
Lorenzo Mazzei et al. / Energy Procedia 126 (201709) 826–833 829
4 Author name / Energy Procedia 00 (2017) 000–000

method for modelling of the soot particle surface area, involved in the calculations of surface-dependent mechanism
of soot formation. The evolution of soot particles’ population can be represented by the population balance equations:
∞ 𝑖𝑖−1 ∞
𝑑𝑑𝑑𝑑1 𝑑𝑑𝑑𝑑𝑑𝑑 1
=− ∑ 𝛽𝛽1,𝑗𝑗 𝑁𝑁1 𝑁𝑁𝑗𝑗 = ∑ 𝛽𝛽𝑗𝑗,𝑖𝑖−𝑗𝑗 𝑁𝑁𝑖𝑖−𝑗𝑗 𝑁𝑁𝑗𝑗 − ∑ 𝛽𝛽𝑖𝑖,𝑗𝑗 𝑁𝑁𝑖𝑖 𝑁𝑁𝑗𝑗
𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑 2
𝑗𝑗=1 𝑗𝑗=1 𝑗𝑗=1

where Ni is the particle number density function that indicates the number of the soot particles of the i-th size class
per unit volume, Bi,j is the collision coefficient between particles of size classes i and j. Except for the first particle
size class, each class is characterized on the right-hand side two source terms, indicating the growth (positive) of Ni
as a result of coagulation between particles of lower size classes and a reduction (negative) due to coagulation between
particles of the current size and any other particles, resulting in formation of particles of higher classes. The solution
of the population balance equations would require to solve an infinite number of particle size classes, which can be
also achieved in a computationally efficient way with the method of moments model. For the particle of the size class
i, its mass mi can be express as mi=im1, where m1 is the mass of a monomer, that is the smallest mass that can be added
or removed from the soot particles, i.e. a single carbon atom. Defined the concentration moment M r of the particle
number density function, two moments are required if the shape of the PSD function in known:
∞ ∞ ∞
𝑀𝑀𝑟𝑟 =∑ 𝑚𝑚𝑖𝑖𝑟𝑟 𝑁𝑁1 𝑟𝑟 = 0,1, … , ∞ 𝑀𝑀0 =∑ 𝑁𝑁𝑖𝑖 =Total particle number density 𝑀𝑀1 =∑ 𝑚𝑚𝑖𝑖 𝑁𝑁𝑖𝑖 =Total mass of the particles
𝑖𝑖=1 𝑖𝑖=1 𝑖𝑖=1

However, additional moments can be added to better predict the shape of the distribution, e.g. M 2 to calculate the
variance and M3 to calculate the skewness.
As moments are affected also by convection and diffusion, a transport equation for the moments can be written:
𝜕𝜕 𝜇𝜇𝑒𝑒𝑒𝑒𝑒𝑒
(𝜌𝜌𝑀𝑀𝑟𝑟 )+𝛻𝛻∙(𝜌𝜌𝑣𝑣⃗𝑀𝑀𝑟𝑟 )=𝛻𝛻∙( 𝛻𝛻𝑀𝑀𝑟𝑟 )+𝑆𝑆𝑟𝑟
𝜕𝜕𝜕𝜕 𝜎𝜎𝑚𝑚𝑚𝑚𝑚𝑚
where Mr is the r-th moment of soot size distribution, σmom the turbulent Prandtl number for moment transport equation
and Sr the source term in the moment transport. It is worth pointing out that the mechanisms involved in the source
term depends on the moment considerer, as the number of particles (represented by M 0) is affected only by nucleation,
coagulation and fragmentation, whereas the source term for soot volume fraction (M1) accounts for nucleation, surface
growth and oxidation. Among the many approaches for the moment closure were presented in literature (an overview
of which is described by Saggese [17]), Fluent implements the interpolative closure, where unknown moments are
interpolated from known ones [18]. Further details concerning the modelling of the formation mechanisms, here
omitted for the sake of brevity, can be found in the ANSYS Fluent Theory Guide [16].

3. Test Case 1: Coelho et al. (2003)

3.1. Description of the test case

The first test case considered consists in a two-dimensional axisymmetric enclosure proposed by Coelho et al. [19].
The domain has a length of L=3.0 m and a radius of R=0.5 m. Walls are black, with a fixed temperature of 300 K.
The enclosure is filled with a mixture at atmospheric pressure, with a volumetric composition of 20% H 2O, 10% CO2
and 70% N2, plus a volumetric fraction of soot equal to 10e-7. The temperature of the medium is 1200 K and 1800 K
in Problem 1 and Problem 2 respectively. The additional Problem 3 consists of a cylinder with L=1.2 m, R=0.3 m and
walls at 800 K, except for the right wall (x=L), which is maintained at 300 K. Soot volume fraction is fixed to 10e-7,
while temperature and molar fractions of H2O and CO2 are given by:

𝑇𝑇(𝑥𝑥, 𝑟𝑟) = 800 + 1200(1 − 𝑟𝑟/𝑅𝑅)(𝑥𝑥/𝐿𝐿) (6)


𝑥𝑥𝐻𝐻2𝑂𝑂 = 0.05[1 − 2(𝑥𝑥/𝐿𝐿 − 0.5)2 ](2 − 𝑟𝑟/𝑅𝑅) (7)
𝑥𝑥𝐶𝐶𝐶𝐶2 = 0.04[1 − 3(𝑥𝑥/𝐿𝐿 − 0.5)2 ](2.5 − 𝑟𝑟/𝑅𝑅) (8)

The ray tracing method with a statistical narrow band spectral model was used to obtain a highly accurate solution
suitable for benchmarking. The spectral absorption coefficient of soot as was modelled as 5.5νfv, where ν is the wave
number and fv represents soot volume fraction.
830 Lorenzo Mazzei et al. / Energy Procedia 126 (201709) 826–833
Author name / Energy Procedia 00 (2017) 000–000 5

3.2. Numerical setup

The calculations were performed on a roughly regular structured grid composed by hexahedrons, consisting of
50x20x20 control volumes, which provide a mesh insensitive solution. The RTE was solved with Discrete Ordinate
model using a 4x4 quadrature. The presence of soot was taken into account by means of a User Defined Function
(UDF), adding its contribution to the value of absorption coefficient calculated by software. A sensitivity analysis was
carried out to assess the impact of different formulations for as. At this purpose, the wide-spread expression reported
in Eq. 3 was implemented, considering C0=5.5 as suggested in Coelho et al. [19]. This formulation was compared
against the model natively implemented in Fluent:

𝑎𝑎𝑠𝑠 = 𝑏𝑏1 𝜌𝜌𝑌𝑌𝑠𝑠 [1 + 𝑏𝑏𝑇𝑇 (𝑇𝑇 − 2000)] (9)

where the coefficients b1=1232.4 m2-kg-1 and bT≈4.8e-4 K-1 were obtained by Sazhin [20] by fitting the equation to
data based on the Taylor-Foster approximation [21] and data based on the approximation made by Smith et al. [4].

3.3. Results

Figure 1 shows the lateral-averaged distribution of the radiative heat flux on the lateral wall for the cases
investigated. The profiles are characterized by a smooth distribution, except in the proximity of left and right walls,
which are significantly colder than gas. It is possible to notice how for this application the gray implementation of the
WSGG model proves a reliable approximation of the problem. It is also worth pointing out that the model proposed
by Sazhin [20] returns a lower radiative heat flux compared to the formulation reported by Modest [6], which for both
problems overestimates the data calculated by Coelho et al. [19], even if the maximum error is however around 10%
(10.49% for Problem 3). Based on these results, and considering the difficulty in distinguishing the role played by the
value estimated for ag, the expression suggested by Modest [6] is retained in the next steps of the present work.

Figure 1: Problem 1 (left), Problem 2 (center) and Problem 3 (right)

4. Test Case 2: Young et al. (1994)

4.1. Description of the test case

The test case chosen to benchmark soot formation modelling was proposed by Young et al. [22]. It represents a
turbulent jet flame burning prevaporized kerosene in a co-flow air stream. The fuel is injected through a 1.5 mm
diameter cylindrical nozzle, whereas the flame is confined by a cylindrical chamber of 155 mm diameter, which allows
optical access and operation at elevated pressure, ranging from 1 to 6.4 bar. Experimental data consist in time-averaged
measurements of soot volume fraction, temperature and mixture fraction. Data were provided along the centerline of
the flame and as radial profiles at two axial locations. The uniqueness of the database makes this test case an interesting
benchmark for modelling and computational prediction [1,13,23].

4.2. Numerical setup


Lorenzo Mazzei et al. / Energy Procedia 126 (201709) 826–833 831
6 Author name / Energy Procedia 00 (2017) 000–000

Exploiting the axisymmetric properties of the test case, the computational domain was reduced to 1 degree, which
allows to obtain a structured hexahedral mesh with 29044 elements. The boundary conditions, reported in Table 1, are
taken from the validated numerical work by Wen et al. [13], as some data such as turbulent quantities were not
provided by Young et al. [22].

Table 1. Boundary conditions for the CFD model

Inlet fuel Inlet air


V (m-s) 22.28 0.234
T (K) 598 288
k (m2-s-2) 0.03 0.03
ε (m -s )
2 -3
0.02 0.02
Wall No slip, smooth, T=300 K
Outlet P=1 bar

The setup in similar to another work carried out with ANSYS Fluent on the same test case [24]. Turbulence was
modelled by means of the standard k-ε model, which, according to a preliminary analysis, provides the best agreement
in terms of mixing. The reactive process in modelled with the steady laminar flamelet model, using the JetSurF Version
1.0 mechanism and generating non-adiabatic diffusive flamelets. Turbulence-chemistry interaction is accounted by
means of PDF integration. The characteristics of the modelling strategy for radiation are already described in the
previous section. In order to assess the impact of the modelling strategy on the solution, four cases are compared:

Table 2: Investigated test matrix

Name Soot model Radiation Fuel


Case 1 Moss-Brookes-Hall No 100% C10H22
Case 2 Moss-Brookes-Hall Yes 100% C10H22
Case 3 Moss-Brookes-Hall Yes 80% decane (C10H22), 20% toluene (C5H6CH3)
Case 4 Method of Moments (e moments) Yes 100% C10H22

4.3. Results

The temperature distribution obtained with the Moss-Brooks-Hall model is reported in Figure 2 to highlight the
main features of the flame under investigation, as well to better understand the position where a quantitative
comparison is carried out (i.e. 100 and 300 mm). As it is possible to notice, the highest temperature is achieved in the
proximity of the inlet, where the stoichiometric conditions are achieved due to the mixing between fuel and air streams.
The axial and radial temperature profiles reported in Figure 3 clearly show that the most fundamental contribution
is given by radiation, especially for x>100 mm. After that in fact the oxidation process leads to a high concentration
of CO2 and H2O that in presence of radiation would reduce flame temperature up to 700 K. Provided that radiation is
accounted, the agreement is very good for the MBH model, while a certain underestimation occurs using the MoM.

Figure 2: Temperature distribution on the symmetry plane (MBH model, w/ radiation, 100% decane)
832 Lorenzo Mazzei et al. / Energy Procedia 126 (201709) 826–833
Author name / Energy Procedia 00 (2017) 000–000 7

Figure 3: Comparison of temperature profiles

The comparison in terms of mixture fraction is reported in Figure 4. As expected the agreement is very good, with
increasing discrepancies among the cases moving downstream. This can be ascribed also to the differences in the
temperature field, which alters the turbulent mixing in radial direction.

Figure 4: Comparison of mixture fraction profiles

Figure 5 show at last the predicted soot volume fraction distributions. It is interesting to observe the overprediction in
the temperature field ascribed to the absence of radiative heat transfer leads to a significant increase in the soot
production (roughly a 2x factor). Once radiation is included the MBH model returns a reasonable agreement in
predicting the maximum value of soot volume fraction, even though the position of such a peak is somewhat shifted
upstream. The prediction is further improved when a more realistic representation of the fuel composition is provided,
i.e. accounting for the aromatic part in kerosene, here modelled with toluene. Such a modification introduces additional
precursors in the flame, thus increasing the soot production. The MoM leads instead to a very poor prediction
characterized by a huge overestimation that justify the excessive flame cooling shown in Figure 3. This could be
ascribed to an approximate estimation of nucleation rate to avoid the specification of the kinetic reactions between
precursors [16]. Such formulation, essentially based on local temperature and concentration of the precursor, exploit
a reduction factor (sticking coefficient) characterized by significant uncertainty.

Figure 5: Comparison of soot volume fraction profiles


Lorenzo Mazzei et al. / Energy Procedia 126 (201709) 826–833 833
8 Author name / Energy Procedia 00 (2017) 000–000

5. Conclusion

This work presents a validation of modelling strategies for the prediction of radiative heat transfer and soot
production. Even a simplified formulation based on a gray participating media is capable of returning acceptable
results for the test cases investigated. Radiative heat transfer is strongly coupled with soot production, altering both
temperature field and soot volume fraction. A secondary effect is given by the appropriate characterization of the fuel,
i.e. accounting for its aromatic content. Overall, the Moss-Brookes-Hall model showed interesting capabilities to
predict the evolution of soot production process in the domain.

Acknowledgements

The authors would like to express their gratitude for the valuable support provided by Dr. Alessandro Innocenti.

References

[1] Bai X, Balthasar M, Mauss F, Fuchs L. Detailed soot modeling in turbulent jet diffusion flames. Twenty-Seventh Symp Int Combust
Combust Inst 1998:1623–30.
[2] Leung K, Lindstedt R, Jones W. A simplified reaction mechanism for soot formation in nonpremixed flames. Combust Flame
1991;87:289–305.
[3] Anderson C, McEnally C, Pfefferle L. Experimental study of naphthalene formation pathways in non-premixed methane flames doped
with alkylbenzenes. Proc Combust Inst 2000;28:2577–83.
[4] Smith T, Shen Z, Friedman J. Evaluation of coefficients for the weighted sum of gray gases model. J Heat Transf 1982:602–8.
[5] Felske J, Tien C. The use of the Milne-Eddington absorption coefficient for radiative heat transfer in combustion systems. ASME J Heat
Transf 1977;99:458–65.
[6] Modest MF. Radiative Heat Transfer. Academic Press; 2013.
[7] Kennedy I. Models of soot formation and oxidation. Prog Energy Combust 1997;23:95–132.
[8] Khan I, Greeves G, Probert D. Prediction of soot and nitric oxide concentrations in diesel engine exhaust. Air Pollut Control Transp
Engines C 1971;142:205–17.
[9] Lindest R. Simplified soot nucleation and surface growth steps for non-premixed flames. Soot Form. Combust., Berlin: Springer-Verlag;
1994, p. 417–39.
[10] Fairweather M, Jones W, Lindstedt R. Predictions of radiative transfer from a turbulent reacting jet in a cross-wind. Combust Flame
1992;89:45–63.
[11] Brookes S, Moss J. Prediction of Soot and Thermal Radiation in Confined Turbulent Jet Diffusion Flames. Combust Flame
1999;116:186–503.
[12] Mueller M, Blanquart G, Pitsch H. A joint volume-surface model of soot aggregation with the method of moments. Proc Combust Inst
2009;32:785–92.
[13] Wen Z, Yun S, Thomson M, Lightstone M. Modeling soot formation in turbulent kerosene/air jet diffusion flames. Combust Flame
2003;135:323–40.
[14] Lee K, Thring M, Beer J. On the rate of combustion of soot in a laminar soot flame. Combust Flame 1962;6:137–45.
[15] Hall R, Smooke M, Colket M. Physical and Chemical Aspects of Combustion. Gordon and Breach; 1997.
[16] ANSYS Fluent Theory Guide, Release 17.0. ANSYS, Inc.; 2016.
[17] Saggese C. Detailed Kinetic Modeling of Soot Formation in Combustion Processes. Politecnico di Milano, 2015.
[18] Frenklach M. Method of moments with interpolative closure. Chem Eng Sci 2002;57:2229–39.
[19] Coelho P, Perez P, El Hafi M. Benchmark numerical solutions for radiative heat transfer in two-dimensional axisymmetric enclosures
with non-gray sooting media. Numer Heat Transf B 2003;43:425–44.
[20] Sazhin S (Fluent E. An approximation for the absorption coefficient of soot in a radiating gas 1994.
[21] Taylor P, Foster P. Some Gray Weighting Coefficients for CO2-H2O-Soot Mixtures. Int J Heat Transf 1974;18:1331–2.
[22] Young K, Stewart C, Moss J. Soot Formation in Turbulent Nonpremixed Kerosine-Air Flames Burning at Elevated Pressure:
Experimental Measurement. Twenty-Seventh Symp Int Combust Combust Inst 1994:609–17.
[23] Balthasar M, Mauss F, Pfitzner M, Mack A. Implementation and Validation of a New Soot Model and Application to Aeroengine
Combustors. ASME J Eng Gas Turb Power 2002;124:66–74.
[24] Rajeshirke P, Nakod P, Yadav R, Orsino S. Parametric Study of Moss-Brookes (MB) and Moss-Brookes-Hall (MBH) Model Constants
for Prediction of Soot Formation in a Turbulent Hydrocarbon Flames 2013:V001T03A009. doi:10.1115/GTINDIA2013-3689.
[25] Khalilian K. Development and Validation of a Partially Coupled, Two-Equation Soot Model for Industrial Applications. Master Thesis.
University of Toronto, 2013.

Vous aimerez peut-être aussi