You are on page 1of 66

Journal Pre-proof

Computational investigation of oxy-combustion of pulverized coal and biomass in a


swirl burner

Ali Cemal Benim, Cansu Deniz Canal, Yakup Erhan Boke

PII: S0360-5442(21)02100-9
DOI: https://doi.org/10.1016/j.energy.2021.121852
Reference: EGY 121852

To appear in: Energy

Received Date: 7 May 2021


Revised Date: 17 August 2021
Accepted Date: 19 August 2021

Please cite this article as: Benim AC, Canal CD, Boke YE, Computational investigation of oxy-
combustion of pulverized coal and biomass in a swirl burner, Energy, https://doi.org/10.1016/
j.energy.2021.121852.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2021 Elsevier Ltd. All rights reserved.


Credit Author Statement:

A. C. Benim: Conceptualization, Methodology, Software, Validation, Formal analysis,

Investigation, Resources, Writing – Original Draft, Writing – Review & Editing, Supervision

C. D. Canal: Validation, Formal analysis, Investigation, Writing – Review & Editing,

Visualisation

E. Boke: Writing – Review & Editing, Resources, Project administration, Funding acquisition

f
r oo
-p
re
lP
na
ur
Jo
COMPUTATIONAL INVESTIGATION OF OXY-COMBUSTION OF PULVERIZED COAL

AND BIOMASS IN A SWIRL BURNER

Ali Cemal Benim1*, Cansu Deniz Canal1,2, Yakup Erhan Boke2

of
1
Center of Flow Simulation, Department of Mechanical Engineering, Duesseldorf,

ro
University of Applied Sciences, Muensterstr. 156, D-40476 Duesseldorf, Germany
2
-p
Faculty of Mechanical Engineering, Istanbul Technical University, Turkey
re
lP

*
Corresponding author: Email: alicemal@prof-benim.com
na
ur

Keywords: Pulverized fuel combustion; oxy-combustion;


Jo

two-phase flow modelling; turbulence modelling; combustion modelling

1
COMPUTATIONAL INVESTIGATION OF OXY-COMBUSTION OF PULVERIZED

COAL AND BIOMASS IN A SWIRL BURNER

A. C. Benim1*, C. D. Canal1,2, Y. E. Boke2

of
1
Center of Flow Simulation, Department of Mechanical Engineering, Duesseldorf,

ro
University of Applied Sciences, Muensterstr. 156, D-40476 Duesseldorf, Germany
1
-p
Faculty of Mechanical Engineering, Istanbul Technical University, Turkey
re
lP

*
Corresponding author: Email: alicemal@prof-benim.com
na
ur

Keywords: Pulverized fuel combustion; oxy-combustion;


Jo

two-phase flow modelling; turbulence modelling; combustion modelling

1
ABSTRACT

Swirling pulverized coal and biomass flames are computationally investigated for oxy-combustion.

The two-phase flow is described by a Eulerian-Eulerian approach. For radiation, the absorption

coefficient is approximated by superposing particle and gas contributions, considering oxy-

combustion conditions for the latter. Turbulence is modelled within a URANS framework, using the

standard k-ε model and Reynolds Stress Model (RSM). It is observed that RSM captures the unsteady

dynamics of the coherent structures, whereas they are not captured by k-ε model. Predicted velocities

of
are compared with measurements. It is observed that the RSM predictions are in a better agreement

ro
with the measurements compared to the k-ε model. The discrepancy between the predictions and

-p
measurements can most clearly be quantified in terms of the peak values of the axial velocity in the
re
forward flow region enveloping the inner recirculation zone. The predictions constantly underpredict
lP

the measurements. On the average, this is about 32% for the RSM and 52% for the k-ε model, for both

flames. The flame is predicted nearly twice as long compared to the coal flame. As means of
na

verification, the coal flame is additionally calculated using a classical Eulerian-Lagrangian two-phase
ur

formulation, leading to quite similar results to the Eulerian-Eulerian formulation.


Jo

2
NOMENCLATURE

aP particle surface area per mixture volume [m-1]

â P reacting particle surface area per mixture volume [m-1]

A pre-exponential factor [corresponding units]

BCA, BK particle burnout parameters [-]

d outer diameter of secondary air nozzle [m]

of
dP particle diameter [m or μm]

ro
Dv(s) s’th percentile of the cumulative particle volume distribution [μm]

-p
E activation energy [J kmol-1] re
f mixture fraction [-]
lP

k turbulence kinetic energy [m2 s-2],

ki pyrolysis or char oxidation rate [corresponding units]


na

Ka absorption coefficient [m-1]


ur

L length, macro length scale [m]


Jo

mi mass fraction of species i [-]

Mi molar mass of species i [kg kmol-1]

pi partial pressure of species i [bar]

r radial coordinate [m]

R universal gas constant [J kmol-1 K-1]

St Stokes number [-]

T static temperature [K]

tF characteristic flow time scale [ms]

3
tP particle relaxation time [ms]

u axial velocity [m s-1]

V velocity magnitude [m s-1]

w swirl velocity [m s-1]

x axial coordinate [m]

Xa radiation model constant [-]

Greek Symbols

γj

of
mass fraction of class j in particle size distibution

ro
Δ LES filter size [m]

-p
ε dissipation rate of k [m2 s-3] re
%ε percentage deviation [-]
lP

θ particle volume fraction [-]

μ viscosity [Pa s]
na

ξC carbon mass fraction of raw fuel [-]


ur

νi,j stoichiometric coefficient of species j in reaction i [-]


Jo

ρ density [kg m-3]

ω specific dissipation rate of k [s-1]

Subscripts

A ash

C char

D diffusional

G gas

i or j species, size class, index

4
max maximum

K kinetical

P particle, pyrolysis

R radiation

T turbulent

0 inlet value

Abbreviations

of
B Biomass

ro
C Coal

-p
CCS Carbon Capturing and Sequestrationre
EE Eulerian-Eulerian
lP

EL Eulerian-Lagrangian

KE k-ε model
na

LES Large Eddy Simulation


ur

NP Number of particle size classes


Jo

RANS Reynolds Averaged Numerical Simulation

RF Raw fuel

RSM Reynolds Stress Model

RWTH Rheinisch-Westfälische Technische Hochschule

URANS Unsteady RANS

VM Volatile matter

WSGGM Weighted Sum of Gray Gases Model

5
1. Introduction

For the generation of power and heat, combustion is being used as the major process to convert the

energy from fossil fuels, since many decades. An overview of different combustion applications can be

found in Lackner et al. [1]. In the last decade, renewable energies are increasingly being used to cover

the energy demand, as it is exemplified e.g. by the recent work of Hussain et al. [2] in the development

of a regional power generation system in Pakistan. In spite of this trend towards renewable energies,

of
combustion continues to play an important role. Suwala [3] pointed out the future importance of fossil

ro
fuels, while meeting more strict environmental requirements, and presented a supply balancing system

-p
for rational planning of the development of the Polish coal industry. Rather recently, Suwala et al. [4]
re
lP

discussed trends in using coal in Poland in international comparison, and emphasized the dominating

role of coal for the near future. Combustion plays an important role, also in the area of renewable
na

energies. This is partly due to the role of hydrogen as a storage medium for renewable power, as it is
ur

manifested e.g. by the study of Shi et al. [5], who studied the hydrogen combustion for a rotary engine.
Jo

The further important reason for the role of combustion in renewable energies is the biomass

utilization. Biomass combustion is increasingly attracting the attention of researchers, as it is

exemplified e.g. by the recent study of Das et al. [6], who experimentally investigated the burning

behavior of different coal-biomass fuel mixtures.

In recent years, a topical issue in the utilization of combustion process for energy conversion has

been the associated CO2 emissions, due to the concerns on global warming and climate change, as

reflected in prospective energy scenarios [7]. This is especially relevant for the utilization of coal. To

address this problem, the oxy-combustion process has been proposed, which has been investigated by

6
many researchers such as Gladysz at al. [8] who presented the thermodynamic analysis of MILD oxy-

combustion, and Qing et al. [9], who performed thermodynamic and economic assessment of oxy-

combustion including the air separation unit. The oxy-combustion process has high potential to reduce

CO2 emissions by enabling efficient carbon capture and sequestration systems (CCS), which has been

the subject of many investigations in its different aspects. Gladysz et al. [10] extended their analysis

on MILD oxy-combustion systems [8] towards CCS, emphasizing the ecological evaluation, while

Zhao et al [11] presented a novel analysis for CCS technology adaption, based on evolutionary game

of
model between stakeholders. In case of combustion with pure oxygen, the exhaust gas consists mainly

ro
of CO2, H2O and excess O2 and has a smaller volume flow rate compared to combustion with air (due

-p
to the missing air nitrogen). This enables a rather easy separation of CO2 by condensation techniques.
re
The missing “dilution effect” of nitrogen leads, however, to too high flame temperatures. To prevent
lP

the latter, in many technical applications, the oxygen stream is diluted by the recycled CO2. Then, the

oxidizer stream consist of oxygen and recycled CO2, which is also the case in the presently considered
na

flames.
ur

Biomass is emerging as a fuel to be used in future energy conversion technologies, not only for its
Jo

being a renewable energy source, but also for its CO2 neutrality. The plants that are the source of CO2

when burnt, capture almost the same amount of CO2 through photosynthesis while growing.

Therefore, biomass combustion is considered to be CO2 neutral. Thus, the renewability and CO2

neutrality make up the main advantages of biomass compared to fossil fuels, particularly to coal. A

flexible option for biomass utilization is its co-firing with coal. Here, its advantages in improving the

CO2 emission performance of the burner can be utilized at least partially, if a full biomass operation is

not possible for any reason (e.g. for large units, where a continuous supply of biomass may not be

guaranteed). In case of biomass, which is CO2-neutral, the application of oxy-combustion with CCS is

7
especially interesting, since a negative CO2 balance can be achieved contributing to the reduction of

CO2 content of the atmosphere.

There are differences in the properties of biomass and coal that put challenges to the development

of new technologies for its efficient firing. On the one hand, there are differences in their composition,

as far as the elementary and the proximate analyses are concerned. On the other hand, there are

differences in their conversion kinetics, and the particle size distributions with important impacts on

the combustion characteristics. Obviously, these aspects need to be addressed adequately in the

of
mathematical modelling.

ro
From the mathematical modelling point of view, combustion of coal and biomass, i.e. solid fuels,

-p
in general, is especially challenging due to the prevailing two-phase flow. Modelling challenges may
re
vary depending on characteristics of the two-phase flow. In fluidized bed combustion with rather large
lP

particles and loadings, special challenges are due to non-negligible particle-particle interactions. The

cyclic nature of coal particles in a circulating fluidized bed was investigated experimentally and
na

mathematically by Gajewski et al. [12] with emphasis on particle spatial temperature distribution. A
ur

detailed experimental analysis of particle velocity distribution in a fluidized bed was presented by
Jo

Leszczynski et al. [13]. Pulverized coal combustion, which is the subject of the present investigation,

is less demanding with respect to particle-particle interactions, due to smaller particle sizes and

loadings.

Pulverized coal combustion was investigated by many researchers in detail over several decades.

An comprehensive overview on simulation procedures for pulverized coal combustion in power

generation was provided by Epple et al.[14]. Benim et al. [15] investigated the particle transport and

separation in a coal pulverizer experimentally and computationally and could obtain a satisfactory

prediction. Kim et al. [16] analyzed MILD combustion of pulverized coal numerically with emphasis

8
on NO formation and showed that NO/char reaction is important to predict NO emissions. A rather

recent computational study of a large-scale, pulverized coal-fired utility boiler with frontal firing was

presented by Madejski [17], where the factors of inefficiency were identified.

In comparison to coal, the combustion of pulverized biomass has not yet been investigated in that

detail. This is especially true for its oxy-combustion. The present work aims to contribute to the

validation of models for the computational modelling of pulverized biomass and coal, under oxy-

combustion conditions. The velocity measurements obtained at the test facility comprising a single

of
swirl burner, located at RWTH Aachen University [18] are taken as the experimental data base, for

ro
validation.

-p
Flames measured at this test facility were computationally investigated by different researchers in
re
the past. Early measurements on this facility for a 100 kWth pulverized coal flame in a CO2/O2
lP

atmosphere were presened by Toporov et al. [18], with emphasis on the adjustment of operation

conditions for stabilizíng the flame for O2 oxidizer content similar to air. Toporov et al. [18] also
na

performed a computational analysis of the 100 kWth oxy-fuel pulverized coal flame, where the k-ε
ur

model was used to model turbulence within a steady-state RANS (Reynolds Averaged Numerical
Jo

Simulation) framework. Large Eddy Simulation (LES) and RANS of the same flame were presented

by Chen and Ghoniem [19]. Compared to RANS, a better agreement of the LES results with the

experiments was observed. However, their LES shall be considered to be a constrained one, since not

the full furnace geometry but a quadrant was considered as the solution domain. The same flame was

investigated by a full LES formulation (considering the whole domain), later on, by Francetti et al.

[20], with emphasis on high-resolution LES modelling (grid with up to 128 million cells). Warzecha

and Boguslawski [21] computationally investigated a 60 kWth pulverized coal flame on the same test

rig, for air- and for oxy-combustion, applying RANS as well as LES, for turbulence modelling. Better

9
overall results were obtained with LES. It was shown that changing the inert gas from N2 to CO2 in the

oxidizer lead to lower flame temperatures. An LES analysis of the same flame was presented by Sadiki

et al. [22], for oxy-combustion, where an emphasis was placed on an accurate modelling of particle-

gas interaction within LES, based on subgrid-scale Stokes numbers. The RANS results were showing

a quite comparable accuracy to LES especially for the velocity field, where LES provided a better

overall accuracy, especially for the temperature and concentration fields [22]. A recent computational

investigation of this flame was also presented by Gaikwad et al. [23], who applied a RANS approach

of
for turbulence modelling using different two-equation turbulence models. NO formation was also

ro
modelled. In the study, a satisfactory overall agreement with the measurements was observed, and a

-p
significant reduction of NO emission by oxy-combustion was manifested [23].
re
The present analysis differs form the above-mentioned investigations in several aspects, which
lP

make up the novelty of the present contribution. The most significant difference lies in the treatment

of the two-phase flow. In all of the above-mentioned investigations, a Eulerian-Lagrangian approach


na

was adopted, where the gas and particulate phases were described by a Eulerian, and a Lagrangian
ur

reference frame, respectively. In the present work, a Eulerian-Eulerian approach is used, where both
Jo

phases are modelled adopting a Eulerian description. This Eulerian-Eulerian model was previously

described and applied to different combustion systems in Benim et al. [24] and Epple et al. [25]. A

further difference exists in the turbulence modelling. In the above-mentioned previous studies, either a

RANS approach with a two-equation turbulence model, or an LES approach was adopted. In the

present work, two-equation as well as Reynolds stress transport turbulence models are used within a

URANS (Unsteady RANS) framework. Furthermore, biomass combustion is investigated, which was

not addressed in the previous studies. Moreover, flames with 40 kWth are considered, where the burner

differs from those of the previous studies in the same furnace by the missing tertiary air injection.

10
2. Modelling

The proposed Eulerian-Eulerian model [24] has been implemented in the finite volume method [26]

based, general-purpose Computational Fluid Dynamics (CFD) software ANSYS Fluent 18.0 [27].

2.1. Outline of the Mathematical Modelling

of
As far as the modellling of the two-phase flow is concerned, gas and solid phases are treated as

ro
interpenetrating continuous phases, both described in a Eulerian frame of reference. The phases are

-p
assumed to be in dynamic and thermal equilibrium. Thus, a single set of momentum and energy
re
transport equations are solved, where the particle phase appears as a “component” of the “mixture”.
lP

The molecular Lewis numbers of all species in the mixture are assumed to be unity [28]. Temperature
na

dependence of the material properties of the mixture such as the specific heat capacity, molecular
ur

thermal conductivity and molecular viscosity are considered by fourth-order polynomials [27,28]. In
Jo

calculating the material properties, a turbulence interaction is neglected, using the average

temperature, neglecting the effect of temperature fluctuations. Similarly, turbulent fluctuations are

neglected in calculating the temperature dependent source term of the radiative equation, as well as

pyrolysis and char oxidation rates. The radiative transport is approximated by the P1 method [27,29].

In many of the above mentioned studies on similar flames, two-equation turbulence modes

were used. In recent years, ω based two-equation turbulence models became popular, as they exhibit

advantages modelling for near-wall turbulence [30,31]. In the present flow, governed by free turbulent

shear layers, the standard k-ε model [32] is preferred as the two-equation turbulence model. On the

11
other hand, it is also known from our previous studies [33-36] that two-equation models do not

perform well in swirling flows and models that can account for the pre-dominant, non-isotropic

turbulence structure of the swirling flow are needed for better accuracy. Thus, the differential

Reynolds Stress Model (RSM) [37] is also employed as a further turbulence model, which solves the

modelled differential transport equation of ε for obtaining the length scale information, in addition to

the solution of modelled differential transport equations for each of the six Reynolds stresses. In both

models, the Reynolds fluxes related with heat and mass transfer are modelled by a gradient-diffusion

of
approximation [27,37], based on turbulent viscosity, assuming a turbulent Prandtl number of 0.9 and a

ro
turbulent Schmidt number of 0.7. In both turbulence models the near-wall turbulence is modelled by

-p
the standard wall-function approach [32], where in case of RSM the wall boundary conditions of
re
Reynolds stresses are obtained based on the k equation [27,37]. In RSM, the quadratic model of
lP

Speziale Sarkar and Gatski [27,38] is used for the pressure-strain correlation. The models are applied

within a URANS framework.


na

The implementation of the present model does not affect the structure of the modelled
ur

transport equations. Thus, the used modelled transport equations are principally those, that are known
Jo

from the literature, with default settings for all model constants [27]. Therefore, they will not be

repeated here, for clarity. Only the aspects will be presented here that are directly related with the

implementation of the model, via material properties and source terms.

In the present Eulerian-Eulerian formularion of the two-phase flow, the mixture density is

calculated from

  1  
p
(1)
mj
RT 
j,G Mj

12
where the summation is to be performed over the gaseous species only (as indicated by the index G).

Besides the definition of the mixture density (Eq. (1)), the implemented model operates on the

absorption coefficient of the radiation transport equation, and the source terms of the species transport

equations as they result from the applied combustion modelling.

2.2. Radiation Properties

of
An ideal forward scattering (idealized Mie scattering) by the particles is assumed [39]. The absorption

ro
coefficient of the mixture (Ka) is calculated from a superposition of the absorption coefficients of the

gas (Ka,G) and particulate (Ka,P) phases as -p


re
lP

K a  K a,G  K a,P (2)


na
ur

Assuming a spherical particle form, the absorption coefficient of a particle cloud can be approximated
Jo

through the specific particle cloud surface area (aP) per mixture volume [39,40]

a
Ka,P  Xa P (3)
4

For the coeffcient Xa , the value of 0.85 is used, which was suggested by Hemsath [40]. The

calculation of the specific particle surface area (aP) is addressed below.

The absorption coefficient of the gas phase is calculated using the so-called “Weighted Sum of

Gray Gases” model (WSGGM) [41]. The mean path length is obtained by a “domain based” approach
13
[27,41]. The WSGGM model of the used software [27] is valid for moderate CO2/H2O partial pressure

ratios [42], which are relevant for combustion with air. For oxy-combustion, different partial pressure

ratios can result, especially for configurations with dry flue gas recycling (like the presently considered

case) where ratios in the order of ten can observed. Thus, for the present analysis, the WSSGM model

of Yin at al. [43] proposed for oxy-combustion conditions is implemented.

2.3. The Pulverized Fuel Combustion Model

of
2.3.1. The Reaction Mechanism

ro
-p
The following four-step global reaction mechanism is employed. In the first reaction the pyrolysis of
re
lP

the raw fuel (RF) is modelled, where the number of particle size classes is given by NP. In the

remaining three reactions, the oxidation of the volatile matter (VM), char (C) and CO are considered.
na
ur

Pyrolysis:
NP
Jo

1,RF RF   1,C j C j  1,VM VM  1,H2O H2O (4)


j1

Volatiles oxidation:
 2,VM VM   2,VM O2   2,CO CO   2,H 2O H 2O   2,N 2 N 2 (5)

Char oxidation:
1
C j  O2  CO (6)
2

CO burn-out:

14
1
CO  O2  CO2 (7)
2

The volatile matter (VM) is represented by hypothetical molecule in the form CαHβOηNζ, which is

designed (adjusting α, β, η, ζ) according to the given fuel composition , assuming a molar mass of 30

kg/kmol. The ash content of the fuel is injected separately, as a further, inert species at the fuel inlet.

2.3.2. The Combustion Model

of
2.3.2.1. Gas Phase Reactions

ro
-p
For the gas-phase reactions (Eqs. (5,7)), the average conversion rate Ri of a species is assumed to be
re
given by
lP
na

R i  min(R K,i ,R T,i ) (8)


ur
Jo

In Eq. (8), RK,i stands for the conversion rate based purely on chemical kinetics, expressed by

Arrhenius expressions [28]. In the volatiles oxidation reaction, (Eq. (5)), for the pre-exponential factor

(A) and the activation energy (E), the values A=2.119.1011 (corresponding SI units), and E=2.027.108

J/kmol, are used in the Arrhenius expression, along with the rate exponents of 0.2 and 1.3 for the fuel

and oxygen, respectively. These values correspond to the default settings of the used software [27] for

the oxidation of a general hydrocarbon fuel to CO and H2O. For the CO burnout reaction (Eq. (7)), the

rate constants suggested by Westbrook and Dryer [44] are employed. The term RT,i in Eq. (8) denotes

15
the mixing rate in the fine scale turbulence eddies. This term is modelled by the Eddy Dissipation

Model (EDM) of Magnussen and Hjertager [45], using the original model constants.

2.3.2.2. Pyrolysis

For determining the rate of the pyrolysis reaction (Eq. (4)), a model based on that of Badzioch and

Hawskley [46] is adopted. Here, conversion rate of raw fuel due to pyrolysis is given by

of
ro
dmRF
 k P mRF (9)
dt

-p
re
where the pyrolysis rate (kP) is modelled by an Arrhenius expression as
lP
na

 E  1
k P  AP exp   P    (10)
 RT  s
ur
Jo

The used rate constants in Eq. (10) will be provided below.

2.3.2.3. Char oxidation

The rate of the char (C) oxidation reaction (Eq. (6)) is modelled based on the assumptions of Field et

al. [47] and Baum and Street [48]. According to the modelling assumptions, the kinetical rate (kK) and

the rate of oxygen diffusion (kD) to the particle surface play a combined role in determining the

effective conversion rate. Assuming a pure CO formation as a result of the heterogeneous reaction of
16
char with oxygen on the particle surface, the effective char oxidation rate for the particle size class j,

per surface are of the coal particle (kC,j) can be expressed by

1  kg 
k C, j   pO 2  2  (11)
 1 1  m s
  
 
 k K k D, j 
 

The kinetic and diffusion rates are given by

of
ro
 E   kg 
k K  A K exp   K   2 

-p
(12)
 RT   m s bar 
 
re
lP

48D0 5 T0.75  kg 
k D, j  10  2  (13)
R T0 d P, j  m s bar 
na

 
ur

The pre-exponential factor AK and the activation energy EK (Eq. 12) are, again, fuel dependent
Jo

parameters. The used values will be provided below. For the constants D0 and T0 that are related with

the diffusion of oxygen through the particle boundary layer, the values proposed by Field et al. [47] are

used (D0=3.49.10-4 m2/s, T0=1600 K). The term dP,j in Eq. (13) is the diameter of the solid fuel particle,

which is assumed to be spherical, in the size class j.

17
2.3.3. Specific Particle Surface Area

For calculating the char conversion rate (Eqs. (11-13)) as well as for calculating the particle phase

absorption coefficient (Eq. (3)), the local specific particle surface area, i.e. particle surface per mixture

volume is required. To this purpose the “shadow” method of Spalding is employed [49]. Here, an

extent of reaction variable BCA is defined as [39,49]

 m  mC  m A

of
BCA  1  C RF (14)

C mRF,0  mA,0 f 

ro
-p
where mRF, mC and mA denote the mass fractions of raw fuel, char and ash, respectively (subscript 0
re
for inlet values), while ξC stands for the carbon mass fraction of raw fuel, as f represent the mixture
lP

fraction. The local mass loss of the raw fuel is modelled by the parameter BK, which is given below
na
ur

mRF  mC  mA
BK  1  (15)
 mRF,0  mA,0  f
Jo

Based on these parameters, the specific particle surface area (aP) can be calculated as [39,49]

1  BCA n B 1  BK 2  mRF  mC  mA 


6
aP  (16)
P,0 d P,0

Where ρS,0 and dP,0 denote the initial values of the particle material density (assumed to be 1400

kg/m3) and the particle diameter, respectively. The parameter nB can take different values to
18
characterizes the burnout behaviour of the particular solid fuel. The value nB=2/3 describes an

idealized case, where the particle burns completely on is outer surface. The value nB=0 corresponds to

the other limiting case, where the particle buns completely from inside (in the pores), which is

assumed in the present model. The “reacting” particle surface ( â P ) can, then, be obtained from [39,49]

1  BCA n B 1  BK 2 mC
6
â P  (17)
P,0 d P,0

of
2.4. Outline of the Numerical Modelling

ro
-p
The coupling of the continuity and momentum equations is treated by a coupled solver [27]. For the
re
spatial discretization, a second order scheme [50] is used for all equations. In time, a bounded second
lP

order implicit discretization is applied [27]. The time step size is chosen in such a manner that the
na

maximum cell convective Courant number [27] does not exceed unity. The applied under-relaxation
ur

factors are 0.75 for velocity and pressure, 0.8 for k and ε, 0.5 for Reynolds stresses, while no under-
Jo

relaxation is applied to the remaining variables. As convergence criteria at each time step, the

threshold values for the scaled residuals of all equations are set to 10-3, except for the energy and

radiative transport equations, while the value 10-6 is set for the latter.

For clarity, a “flowchart” of the model is sketched in Fig. 1. The solution sequence of the

transport equations is illustrated on the left side of the figure. On the right side, the main inputs by the

presently proposed model and their interaction with the transport equations are schematically outlined.

19
of
ro
-p
re
lP
na

Fig. 1. The logical scheme of the model.


ur

3. Furnace and burner geometry, operating conditions, fuel properties


Jo

As mentioned above, the considered flames were obtained and measured at a test rig of RWTH

Aachen University [51]. Sketches of the test rig and the swirl burner are presented in Fig. 2.

20
(b)

of
ro
-p
(a) re
lP

Fig. 2. Sketch of test rig of RWTH Aachen University [51],

(a) combustion chamber, (b) swirl burner.


na
ur

Within the present scope, two pulverized solid fuels, namely coal (pre-dried lignite) and biomass
Jo

(torrified biomass) are considered as fuels, which will, later on, be abbreviated as (C) and (B),

respectively.

In both cases, the oxidizing agent is a mixture of O2 and CO2. Both flames have a thermal power

of 40 kWth. The operating conditions for both flames [51] are summarized in Table 1.

21
Table 1. Operating conditions [51]

Parameter C B

Fuel type Coal Biomass

Fuel mass flow rate (injected with primary stream) [kg/h] 6.7 7.0

Primary gas stream flow rate [mn3/h] 9.4 9.4

O2/CO2 primary gas stream [vol%] 24.4/75.6 24.4/75.6

Temperature of primary gas stream [°C] 25 25

of
Secondary gas stream flow rate [mn3/h] 8.8 8.5

ro
O2/CO2 secondary gas stream [vol%] 27/73 27/73

-p
Temperature of secondary gas stream re [°C] 40 40

Staging gas stream flow rate [mn3/h] 23.1 22.7


lP

O2/CO2 staging gas stream [vol%] 27/73 27/73


na

Temperature of staging gas stream [°C] 900 900


ur
Jo

A swirling flame is achieved by swirling the secondary gas stream. The geometric swirl number of the

secondary air was given as 0.958 [51], whereas in the computational previous studies [19], a higher

effective swirl number of 1.06 was reported, based on the detailed analysis of the burner flow.

The ultimate and proximate analysis [51] of the fuels are given in Table 2. Please note that the

sulphur content is neglected in the present calculations.

22
Table 2. Measured [51] ultimate and proximate analysis

Fuel Coal (C) Biomass (B)

Component AR DAF AR DAF

Carbon [w-%] 58.21 68.20 55.15 57.69

Hydrogen [w-%] 4.00 4.69 5.52 5.77

Oxygen [w-%] 21.94 25.71 34.65 36.24

Nitrogen [w-%] 0.83 0.97 0.25 0.26

of
Sulphur [w-%] 0.37 0.43 0.03 0.04

ro
Water [w-%] 9.13 - 2.26 -

-p
Ash [w-%] 5.52 re - 2.14 -

Volatiles [%] 44.99 52.71 68.71 71.84


lP

LHV [MJ/kg] 21.423 25.361 20.525 21.527

HHV [MJ/kg] 22.512 26.376 21.778 22.780


na
ur
Jo

The information on the particle size distribution provided in Ref. [51] is presented in Table 3.

Table 3. Measured [51] particle size distribution

Dv(10) [μm] Dv(50) [μm] Dv(90) [μm]

Coal (C) 5 29 132

Biomass (B) 5 18 46

23
4. Solid fuel modelling: particle size distribution, kinetics data

The provided particle size distribution information in Ref. [51] (Table 3) does not have a high

resolution. Based on the given information, and also relying on the particle size distribution published

for a similar case [52], continous Rosin-Rammler distributions [53] are first derived. They are, then,

discretely approximated by five particle size classes, for each fuel. The resulting particle size

distributions that are used in the present calculations are presented in Table 4.

of
Table 4. The assumed particle size distributions.

ro
Mass fraction (γj) [%]

-p
0.1 0.2 re 0.4 0.2 0.1

Particle size (dP,j) [μm]


lP

Coal (C) 3 7 18 55 178

Biomass (B) 3 6 16 32 53
na
ur
Jo

A parameter with a strong effect on the prediction of pulverized fuel combustion is the pyrolysis rate.

Unfortunately, it is largely dependent on the fuel type and its determination is normally afflicted with a

rather high amount of uncertainty. In the presently applied single-rate pyrolysis represented by an

Arrhenius rate expression (Eq. 10), the pre-exponential factor (AP) and the activation energy (EP) need

to be defined. For coal, the values are used, which were used in the computational work of Sadiki et

al. [22] (LES) and Gaikwad et al. [23] (URANS), for the same coal (burnt by a different, 60 kWth

burner). For the rate constants of biomass, the empirical rate constans suggested by Tolvanen et al.

24
[54] for torrified wood are used, which were obtained in a drop tube reactor. The presently employed

pyrolysis rate constants are summarized in Table 5.

Table 5. Pyrolysis rate constants.

Fuel AP [1/s] EP [J/kmol]

Coal (C) [22,23] 2.0.105 4.8.107

Biomass (B) [54] 1.925.104 5.678.107

of
ro
It shall be noted that the values for the coal, in Table 5, are not necessarily in agreement with

-p
the more commonly used values, such as EP=7.4.107 J/kmol [14,16,24,25,40,46-48]. Still, these values
re
(Table 5) are used, in the present work, for a better compatibility with the previous related work
lP

[22,23] using the same coal (also considering the fact that the researchers involved in the

measurements were also co-authors of the mentioned previous work [22]). By this choice (with a
na

rather low activation energy for coal, EP=4.8.107 J/kmol) the activation energy of the biomass,
ur

EP=5.678.107 J/kmol, borrowed from Ref. [54], turns out to be larger than that of coal.
Jo

For char oxidation (Eq. (12)), rate constants suggested by Smoot and Pratt [55] are quite

commonly used in case of coal combustion. In the present work, the same rate constants are used for

coal and biomass combustion, both, due to the lack of more detailed information. In the present model,

the char conversion rate is not controlled purely by kinetics, but by a combined effect of diffusion and

kinetics (Eq. 11) . Due to this combined control, this uncertainty in the kinetical rate constants is

expected to play less a role. The used values are: AK = 204.0 1/s, EK = 79.4.106 J/kmol.

25
5. Solution domain, boundary conditions

The geometry and the boundary conditions of the problem is axially symmetric. This means that the

converged, time-averaged can only be also axi-symmetric, implying that a RANS calculation in a 2D-

axisymmetric domain could be adequate. However, our previous experience [33-36] indicates that the

unsteadiness of coherent structures (that are important in turbulent swirling flows) cannot be

adequately be represented by RANS, necessitating an unsteady approach, such as URANS (or LES for

of
a finer resolution). Thus, a URANS formulation is adopted. Since the unsteadiness of the coherent

ro
structures are closely related with three-dimensionality, the solution domain is also defined to be

-p
three-dimensional in the present study. re
The experimental furnace is quite long compared to burner diameter (Fig. 2). It can, however be
lP

observed that the resulting flames [51] are confined to a rather short region downstream the burner.

Thus, not the whole furnace geometry, but a region upto 600 mm downstream the burner exit plane is
na

covered by the solution domain. It is, of course, ensured that this does not cause any inaccuracy in the
ur

results, which will be demonstrated below. A domain length of 600 mm was used also by Sadiki et al.
Jo

[22], in their LES study of pulverized coal flames (by a 60 kWth burner) in the same furnace, which

supports the present definition of the domain length. A further issue regarding the domain definition in

swirling flows is the definition of the outlet boundary. Special care is needed in the formulation of

outlet boundary conditions in swirling flows, since they can have strong influence on the upstream

flow, especially in case of sub-critical flow [56]. From the practical point of view, in case of a

straightforward, classical “axial outflow” boundary one can easily run into convergence problems,

since the inner recirculation zone generated by vortex breakdown may reach the outlet boundary and

cause backflow into the domain. To minimize the inaccuracies and complications in this respect, we

26
have found, in our previous studies [33-36] that it is most convenient to replace the axial outflow

boundary by a radial outlet, without affecting the results in the region of interest. A sketch of the

solution domain and the boundary types are presented in Fig. 3.

of
ro
-p
re
lP
na

Fig. 3. Solution domain, boundaries.


ur
Jo

At all inlets, top-hat velocity profiles are prescribed for all convective-diffusively transported variables

as implied by with the operating conditions and fuel properties (Tables 1, 2, 4). For turbulence

quantities, the inlet conditions are derived from an assumed turbulent intensity of 4% and a length

scale based on the hydraulic diameter. An isotropic turbulence state is assumed at the inlets. At the

outlet, the static pressure is prescribed along with zero normal gradient for the convective-diffusively

transported variables. At the walls, the no-slip boundary condition apply for the momentum and

turbulence transport equations (amended by the wall functions), while zero normal gradient condition

is used for the species transport equations. For the energy equation, a temperature of 800°C is applied

27
on the cooled furnace walls, in accordance with the experiments [51]. For the radiative transport

equation, the wall emissivity is set to the value of 0.7, while inlet and outlet boundaries are assumed to

be black surfaces.

Axial variations of the predicted time-avergaged axial centerline velocity for different domain

sizes, namely for assumed furnace lengths of 600 mm and 2100 mm are displayed in Fig. 4. These

calculations have been performed for isothermal, air flow using the k-ε model. In the calculation with

the long furnace (2100 mm), a “classical” “axial outlet” boundary is used, whereas the proposed

“radial outlet” boundary (Fig. 4) is employed in the calculation with the short furnace (600 mm). The

of
ro
whole range of the measuring stations, where experimental data was obtained [51] is also indicated in

-p
the figure. One can see in Fig. 4 that both curves agree very well, especially in the range of
re
experimentally investigated region.
lP
na
ur
Jo

Fig. 4. Axial variation of time-averaged centerline velocity for different furnace lengths and outlet

boundaries for isothermal air flow (k-ε model).

28
6. Grid

A block structured strategy based on hexahedral cells is used in grid generation. A grid independence

study is performed based on the isothermal air flow, using the k-ε model, assuming that this grid

would provide sufficient grid independence also for the cases to be considered. The total number of

cells have been (rounded values): 3.9.105 (Grid 1), 5.6.105 (Grid 2), 7.5.105 (Grid 3), 1.0.106 (Grid 4),

1.5.106 (Grid 5). The radial variations of the time-averaged swirl velocity (w) at the axial position

of
ro
x/d=0.3 predicted by the five grids are displayed in Fig. 5.

-p
re
lP
na
ur
Jo

Fig. 5. Radial variation of time-averaged swirl velocity at x/d = 0.3, for different grids, for isothermal

air flow (k-ε model).

29
As can be seen in Fig. 5, the results of the Grids 3, 4, 5 are quite close. The Grid 4 with about 1.0.106

cells is used in the calculations. Views of the grid are presented in Fig. 6.

of
ro
(a)
-p (b)
re
lP
na
ur
Jo

(c)
(d)

Fig. 6. Views of the grid, (a) surface grid furnace, (b) surface grid burner, (c) grid in longitudinal

section, (b) grid in furnace cross-section.

30
In the previoius work of Sadiki et al. [22], where a different burner (60 kWth) in the same furnace,

with the same assumed length (600 mm) was investigated by LES, the employed grid had also about

1.0.106. This supports the adequacy of the present grid resolution, since the presently applied URANS

approach is principally less demanding compared to LES. An LES grid resolution criteria is also

applied to the present grid, as a further check of the grid resolution. According to Gant et al. [57], for

an adequate grid resolution for LES, the condition L/Δ > 12 should be fulfilled, where L and Δ denote

the turbulence macro length scale, and the cell size, respectively. Obtaining L from L=k1.5/ε, and Δ

from the third root of the local cell volume, the calculated L/Δ ratio for the case of coal combustion,

of
ro
using the k-ε model, is presented in Fig. 7, for a middle plane through the furnace.

-p
re
lP
na
ur
Jo

Fig. 7. Predicted L/Δ distribution along a mid-plane through furnace (k-ε model, coal combustion).

One can see that the present grid is fairly fine also in this perspective, as considerably large regions

satisfy the above-mentioned LES grid resolution criterion. However, one can also see that the grid is

31
still not sufficiently fine for LES, since in the especially critical regions, where the swirling jet

expands into the furnace, the grid resolution criterion is not fulfilled satisfactorily (Fig. 7).

7. Results

As result of the URANS analysis, it is observed that the calculations based on the k-ε turbulence

of
model could not capture any flow unsteadiness and converged to a steady-state solution. The Reynolds

ro
Stress Model (RSM) calculations could capture low frequency unsteadiness of coherent structures.

-p
This is in agreement with the previous experience [33-36] of the present authors in swirling flames.
re
At this stage, the adequacy of the present Eulerian-Eulerian approach for the presently
lP

considered flames shall be assessed by a dimensional analysis. A basic assumption of the present

model is a dynamic and thermal equilibrium between the phases. This assumption can be assessed
na

with the help of the Stokes number that is defined as St=tP/tF [60], where tP stands for the particle
ur

relaxation time (particle response time), and tF represents a characteristic flow time scale. For small
Jo

Stokes numbers (St<1), an equilibrium between the phases can be assumed to be a reasonable

assumption. The particle relaxation time, defined as tP=ρPdP2/18μ [60], depends on the particle size

(dP). To achieve an overall assessment for the given size distributions (Table 4), the mass mean

diameters are taken to represent the distributions cumulatively, i.e. on the average, leading to tP≈0.65

ms and tP≈0.18 ms for coal and biomass, respectively. A characteristic flow time scale can be obtained

from the velocity magnitude, and the burner cone diameter at the inlet, leading to tF≈3.33 ms. These

values lead to Stokes numbers of St≈0.20 for coal and St≈0.05, which imply that the present approach

may be considered to be adequate.

32
7.1. Field distributions

The predicted iso-surface of the Q-criterion [58] by the RSM, at a time step, for the value of 1000 s-2

is plotted in Fig. 8 for the coal and biomass flames, where the iso-surface is colored by the velocity

magnitude.

of
ro
-p
re
lP
na

(a)
ur
Jo

(b)
Fig. 8. Predicted iso-surface of Q-criterion (for iso-value of 1000 s-2) at a time step, predicted by
RSM, colored by velocity magnitude, (a) coal flame, (b) biomass flame.
33
The observed structures of the Q-criterion iso-surface manifest the existence of unsteady, three-

dimensional coherent structures, and the existence of the so-called “precessing vortex core” [59], in

both flames. As mentioned before, such unsteady, three-dimensional vortex structures could not be

captured by the k-ε model.

The predicted distributions of velocity magnitude for the coal and biomass flames by the k-ε

model and RSM (at an instant of time, and time averaged) in a middle plane through the combustor

are displayed in Fig. 9. One can see that the k-ε model and RSM predictions have qualitative

of
ro
similarities. However, they differ quantitatively. The difference between the k-ε model and RSM

-p
predictions are larger for the biomass flame compared to the coal flame (Fig. 9).
re
The predicted distributions of temperature for the coal and biomass flames by the k-ε model and
lP

RSM (at an instant of time, and time averaged) in a middle plane through the combustor are presented

in Fig. 10. In all predictions, one can observe that the coal flame is, in general, shorter than the
na

biomass flame, with slightly higher local temperatures. According to the RSM predictions of the
ur

temperature field, the biomass flame is twice as long compared to the coal flame, and exhibits the
Jo

character of a lifted flame (Fig. 10). The main reason for this comparative behaviour is expected to be

the difference in the pyrolysis rates. As can be seen in Table 5, the used pyrolysis rate constants imply

a much faster pyrolysis of coal compared to biomass, which can be the main reason for this difference

in the predicted flame lengths.

Distributions of the mass fractions of three (Table 4), namely the first, third and fifth, char size

classes (mC1, mC3, mC5) predicted by the RSM, are presented in Fig. 11, in a middle plane through the

combustor, at an instant of time. The above-mentioned difference in the pyrolysis rates can clearly be

observed in distributions of the char size class mass fractions. For coal, the pyrolysis starts much

34
earlier and the char starts to get released quite quickly downstream the primary injection resulting in

high char mass fractions even in the burner cone. For the biomass flame, due to the assumed pyrolysis

model implying lower rates, the release of char starts much later, at a quite downstream position from

the burner, leading to a flame with a rather lifted character. Although the char size classes are released

quite late at quite downstream positions in the biomass flame, char burnout takes places more quickly

compared to coal. This is due to the smaller particle sizes of the biomass size class distribution (Table

4).

of
ro
-p
re
lP
na
ur
Jo

35
(a) (d)

of
ro
-p
re
lP
na

(b) (e)
ur
Jo

(c) (f)

Fig. 9. Predicted distribution of velocity magnitude in a middle plane through combustor, for coal

(left) and biomass (right), (a) C-KE, (b) C-RSM at time step, (c) C-RSM time averaged, (d) B-KE, (e)

B-RSM at a time step, (f) B-RSM time averaged.


36
(a) (d)

of
ro
-p
re
lP
na

(b) (e)
ur
Jo

(c) (f)

Fig. 10. Predicted distribution of temperature in a middle plane through combustor, for coal (left) and

biomass (right), (a) C-KE, (b) C-RSM at time step, (c) C-RSM time averaged, (d) B-KE, (e) B-RSM

at a time step, (f) B-RSM time averaged.


37
(a) (d)

of
ro
-p
re
lP
na

(b) (e)
ur
Jo

(c) (f)

Fig. 11. Char size class mass fractions at a time step in middle plane through combustor, by RSM, for

coal (left) and biomass (right), (a) C, mC1, (b) C, mC3, (c) C, mC5, (d) B, mC1, (e) B, mC3, (f) B, mC5.

38
7.2. Line Plots

Predictions are compared with the measurements in terms of radial profiles at different axial positions.

The predicted radial profiles of the time averaged axial (u) and swirl (w) velocity components are

compared with the measurements [51] at four axial positions in Fig. 12 and Fig. 13, respectively (d:

outer diameter of secondary air nozzle). Qualitatively, the predictions and measurements indicate

similar general flow patterns. Along the combustion chamber and in the nearfield of the burner, an

of
internal recirculation zone (IRZ) is formed, which is typical for swirl stabilized flames. Experiments

ro
indicate a similar recirculation intensity for coal and biomass flames. This region is overlapping with

-p
the main vortex and exhibits substantial negative axial velocities. For both flames, both turbulence
re
models predict a longer and broader IRZ, compared to the measurements, which can be deduced by
lP

inspecting the axial velocity profiled (Figs. 12a, 13a).

The predicted maximum negative velocities along the centerline show a fair agreement with the
na

experiments for the the initial axial stations (x/d=0.3, 0.5), but the prediction quality deteriorates
ur

further downstream. For both flames, the axial and radial extensions of the IRZ is overpredicted by the
Jo

calculations. The experimentally obtained maximum axial velocities on the edge of the IRZ are quite

high and underpredicted by the calculations at all axial locations, for both flames. In that respect, one

can see that the predictions by the RSM is clearly better than those of the k-ε model. In the

measurements, a secondary local maximum is observed for the axial velocity at a radial position about

0.06 m, for axial stations near the burner (x/d=0.3, 0.5). This could not be predicted.

The measured swirl velocity profiles also exhibit two local maxima at axial positions near the

burner (x/d=0.3, 0.5). This is qualitatively predicted by the RSM results. However, quantitatively, the

peak values of the swirl velocity are underpredicted, although the predicted size of the vortex core by

39
the RSM is in a fair agreement with the experiments. Also for the swirl velocity, a much better

agreement of the RSM results with the experiments can be observed, compared to the k-ε model.

However, the RSM prediction quality still leaves much to be desired, quantitatively.

In comparing the predicted velocity profiles with the measurements, the underprediction of the

peak axial velocity on the edge of the IRZ can be recognized as a typical behavior, which is observed

for both turbulence models, for both flames. The degree of underprediction is about 32% for RSM and

52% for the k-ε model, on the average, for both flames.

of
In the publication of Zabrodiec et al. [51], which is taken as the basis for the considered test

ro
case and the validation data, no temperature measurements were provided. Therefore, a comparison of

-p
the predicted temperature fields with the measured temperatures has not been possible. The predicted
re
radial profiles of the time averaged temperature (T) by the k-ε model and RSM are compared with
lP

each other in Fig. 14, at the same four axial positions considered previously for the velocity

components (Figs. 12, 13). Higher temperatures for the coal flame compared to the biomass flame are
na

observed for the downstream axial stations (x/d=1.0, x/d=2.0, Figs. 14c,d,g,h) reflecting the lifted
ur

nature of the biomass flame, which was already discussed in relationship with the temperature contour
Jo

plots (Fig. 10).

For both flames, the temperature curves obtained by the k-ε model and RSM are rather close to

each other at upstream axial stations (x/d=0.3, x/d=0.5) and start to show larger deviations further

downstream (x/d=1.0, x/d=2.0). The burner inlet conditions (Table 1) imply a quite under-

stoichiometric mixing (equivalence ratio of about 1.7). In the burner, the reaction progress is quite low

and the temperatures profiles close to burner outlet are dominated by the recirculating gases. This can

be seen to explain the comparably small difference between the temperature curves of both turbulence

models, at the upstream axial stations (x/d=0.3, x/d=0.5, Fig. 14). Mixing of additional oxidizer

40
through the staging stream (overall equivalence ratio is about 0.7) occurs in the shear layers emerging

from the burner edge. This leads to a further reaction progress with an accompanying temperature

increase in the shear layers, in the downstream, where the predictions of the two turbulence models

become more apparent (x/d=1.0, x/d=2.0; Fig. 14).

of
ro
-p
re
lP
na
ur
Jo

41
(a) (e)

of
ro
(b) (f)
-p
re
lP
na
ur
Jo

(c) (g)

(d) (h)

Fig. 12. Measured [51] and predicted radial profiles of time averaged axial (left) and swirl (right)
velocities for coal flame at four axial stations (a) u at x/d=0.3, (b) u at x/d=0.5, (c) u at x/d=1.0, (d) u
at x/d=2.0, (e) w at x/d=0.3, (f) w at x/d=0.5, (g) w at x/d=1.0, (h) w at x/d=2.0.
42
(a) (e)

of
ro
(b) (f)
-p
re
lP
na
ur
Jo

(c) (g)

(d) (h)

Fig. 13. Measured [51] and predicted radial profiles of time averaged axial (left) and swirl (right)
velocities for biomass flame at four axial stations, (a) u at x/d=0.3, (b) u at x/d=0.5, (c) u at x/d=1.0,
(d) u at x/d=2.0, (e) w at x/d=0.3, (f) w at x/d=0.5, (g) w at x/d=1.0, (h) w at x/d=2.0.
43
(a) (e)

of
ro
(b)
-p (f)
re
lP
na
ur
Jo

(c) (g)

(d) (h)
Fig. 14. Predicted radial profiles of time averaged temperatures for coal (left) and biomass (right) at
four axial stations, (a) C, x/d=0.3, (b) C, x/d=0.5, (c) C, x/d=1.0, (d) C, x/d=2.0, (e) B, x/d=0.3, (f) B,
x/d=0.5, (g) B, x/d=1.0, (h) B, x/d=2.0.
44
7.3. A Comparison with an Eulerian-Lagrangian Based Formulation

As already discussed above, the commonly employed computational models for pulverized fuel

combustion are based on the Eulerian-Lagrangian (EL) formulation of the two-phase flow, where the

gas phase is described by a Eulerian, and the particle phase by a Lagrangian frame of reference. The

main distinguishing feature of the present model is an Eulerian-Eulerian (EE) formulation of the two-

phase flow, where an dynamic and thermal equilibrium between the phases is additionally assumed. In

of
the present section, a comparison between the present EE formulation with the classical/standard EL

ro
formulation is aimed, for the considered cases. The dimensional analysis based on the Stokes number

-p
presented above indicates that the adequacy of the present EE formulation is more an issue for the coal
re
flame. Therefore, the comparison is done only for the coal flame. Since a comparison between the
lP

two-phase formulations is aimed, this being not directly related with turbulence modelling, the
na

comparison is done using the k-ε model only.


ur

To this purpose, the present coal flame is simulated using the standard pulverized coal
Jo

combustion model of ANSYS Fluent, based on the standard/classical EL formulation. In these EL

calculations, the turbulence model (k-ε), boundary conditions and grids are the same as those used in

the present EE calculations. An exception lies in the radiation properties. In both calculations, the P1

radiation model based on WSGGM is used. However, in the present EE calculations, the WSGGM of

Yin et al. [43] is implemented and used, which was developed for oxy-combustion conditions,

whereas in the EL calculations, the default WSGGM model [42] suitable for air combustion is used.

Beyond this aspect, the differences in the results can be traced back to the two-phase flow modelling.

45
In the EL simulation of the pulverized coal combustion, where the trajectories of individual

particles are calculated within a Lagrangian formulation, the default settings of ANSYS Fluent are

used, for all related model parameters. For keeping similarity to the present EE calculation, the same

particle size distribution based on five size classes is employed. A special feature of the Lagrangian

modelling is the consideration of the turbulent particle dispersion by the so-called “random walk”

model [61]. Here, particle trajectories are calculated by considering the turbulent velocity fluctuations.

of
To obtain a stable average, a sufficiently large number of trajectories need to be calculated. Each such

ro
calculation is termed as “trial”. In the common practice [24,25], at least five trials are mostly applied,

-p
where a higher number [24] may be preferred for better accuracy. In the present comparison, five trials
re
are used.
lP

Calculated particle trajectories by the EL formulation are displayed in Fig. 15, for the

pulverized coal flame. The trajectories are colored by the particle temperature. For clarity, only a small
na

number of trajectories are displayed, i.e. only 750 out of 15,000 (The number of inlet cells is 600.
ur

With 5 particle size classes, it makes 3,000 particles, and with 5 trials, the total number becomes
Jo

15,000).

The spread of the fuel jet under the influence of the swirling motion, the turbulent dispersion

of the particles, as well as the heating up of the particles towards the flame zone can be recognized in

Fig. 15.

The temperature distribution in a middle plane through the combustor as predicted by the

present EE model is compared with that of the EL calculation, for the coal flame, in Fig. 16 (The EL

prediction is presented in the upper half and the present EE prediction is presented in the lower half of

the figure).

46
of
ro
-p
Fig. 15. Sample particle trajectories in burner nearfield predicted by the EL model, for the pulverized
re
coal flame (C) (k-ε turbulence model).
lP
na
ur
Jo

Fig. 16. Temperature fields in middle plane through combustor predicted by the EL (upper half) and
the EE (lower half), for pulverized coal flame (C) (k-ε turbulence model).
47
One can see that the predicted temperature fields are, in general, quite similar, although there

are differences. The flame position on the furnace axis is similarly predicted. According to the EL

predictions, the ignition seems to occur slightly earlier, and the flame brush is thinner compared to EE

(Fig. 16). It shall also be recalled that the radiation properties are calculated differently in the both

calculations, as already stated above. Thus, the predicted temperature fields by EE and EL can be

assumed to show a fair correspondence (Fig. 16).

of
The radial profiles of the axial (u) and swirl (w) velocity components predicted by the presently

ro
proposed EE formulation and the standard EL formulation, for the coal flame, using the k-ε model are

-p
compared in Fig. 17. Alhough the current purpose is the comparison between the EE and EL models,
re
the experimental curves are kept in the figures for a better orientation. One can see that the EE and EL
lP

results are quite close to each other, which supports the adequacy of the presently advocated EE

formulation.
na
ur
Jo

48
(a) (e)

of
ro
(b)
-p(f)
re
lP
na
ur
Jo

(c) (g)

(d) (h)
Fig. 17. Comparison of the EE prediction (PRED (EE)) with the EL prediction (PRED (EL)) for the
radial profiles of axial (left) and swirl (right) velocities for coal flame at four axial stations (a) u at
x/d=0.3, (b) u at x/d=0.5, (c) u at x/d=1.0, (d) u at x/d=2.0, (e) w at x/d=0.3, (f) w at x/d=0.5, (g) w at
x/d=1.0, (h) w at x/d=2.0.

49
The appropriateness of the present EE model for the present cases can further be assessed, by

comparing the gas velocities and temperatures with those of the particles along the particle trajectories

calculated by the EL model. Small differences between the gas and particle values would imply the

validity of the present equilibrium assumption and suggest comparability of the EE and EL

formulations. To this purpose, five sample particle trajectories are calculated (EL), each for a size class

of the present size distribution (Table 5). Doing so, the random-walk model is turned-off, for

of
simplicity. Average gas residence time in the furnace is approx. 1.2 s. Gas and particle velocities and

ro
temperatures along particle trajectories are monitored for a period of 0.6 s, which approximately

-p
corresponds to the first half of the furnace that accommodates the flame zone, approximately. The
re
velocity magnitudes and temperatures for the gas and particles along trajectories of the particles of the
lP

five size classes are presented in Fig. 18.

One can see that the results for the gas and particles are practically identical for the two
na

smallest particle size classes (dP=3 μm, 7 μm) and very close for the medium size class (dP=18 μm).
ur

These three size classes make up 70% of the total particle mass (Table 4). For the fourth size class
Jo

(dP=55 μm), which represents a mass fraction of 20% (Table 4), rather small differences are observed

for a short period. Substantial differences are observed for the fifth, i.e. for the largest size class

(dP=178 μm). Although these differences are rather large, it shall be recalled that this size class

(dP=178 μm) makes up only 10% of the total particle mass. Based on the curves shown in Fig. 18, an

average percentage deviation (averaged over the time period of τ=0.6 s) between the gas

and particle predictions can be calculated for each size class (j), for the velocity magnitude (%εV,j) and

for temperature (%εT,j) as %εV,J=100x(1/τ)∫t|VG-(t)-VP(t)|dt/VP,0 and %εT,j=100x∫t|(TG(t)-

TP(t))|dt/(TP,max-TP,0). These can be weighted by the mass fraction of each size class (γj, Table 4) to

50
obtain a cumulative, i.e. overall assessment for the whole size class distribution, for the velocity (%εV)

and temperature (%εT), as %εV=Σj(γj)(%εV,j) and %εT=Σj(γj)(%εT,j). The cumulative percentage

deviations between the gas and particle results for the velocity magnitude and temperature result as

%εV=7 and %εT=4. These values are rather small and confirm the observed similarity between the EE

and EL predictions (Figs. 16, 17) as well as the Stokes number based dimensional analysis presented at

the beginning of the results section. This comparison of gas and particle velocities and temperatures

along particle trajectories (Fig. 18) with obtained cumulative percentage deviations (%εV=7, %εT=4)

of
supports the adequacy of the present EE modelling approach for the presently considered flames.

ro
-p
re
lP
na
ur
Jo

51
(a) (f)

of
(b) (g)

ro
-p
re
lP

(c) (h)
na
ur
Jo

(d) (i)

(e) (j)
Fig. 18. Comparsion of gas and particle speeds (left) and temperatures (right) along trajectories of five
particles with different size, predicted by the EL model for the coal flame, based on the k-ε model,
(a) V, dP=3 μm, (b) V, dP=7 μm, (c) V, dP=18 μm, (d) V, dP=55 μm, (e) V, dP=178 μm,
(f) T, dP=3 μm, (g) T, dP=7 μm, (h) T, dP=18 μm, (i) T, dP=55 μm, (j) T, dP=178 μm.
52
Finally, the significance of the presently advocated EE model for the prediction of the

pulverized fuel flames, in general, shall be underlined. The present EE model does not necessarily

claim a better accuracy compared to the EL model. The EL formulation has the potential of being more

accurate, since no equilibrium between the phases is assumed. The main advantage of the EE

formulation lies in its lower CPU demand. In many situations, the Lagrangian particle trajectory

calculation becomes quite time consuming. In the present EL calculations, a rather low number of

of
trials, i.e. 5, is used to model turbulent particle dispersion. Still, one iteration over the gas and particle

ro
phases in the EL simulation requires nearly four time more CPU time compared to one iteration of the

-p
EE simulation. This indicates the comparatively high CPU demand of the EL model. In this respect, a
re
very important point to note is the following: The computational effort related with the particle
lP

calculation in the EL formulation increases in proportion with the number of inlets, since for any

additional inlet, additional particles need to be injected. In the EE formulation, where differential field
na

equations are solved for the particle phase, the number of inlets does not have any influence on the
ur

number of equations and, thus, on the computational effort. Therefore, the EE formulation, which has
Jo

now been assessed for the present laboratory burner, is especially suitable and attractive alternative for

the simulation of large utility boilers, which normally exhibit a quite large number of burner inlets.

53
8. Conclusions

Pulverized coal (lignite) and biomass (torrified biomass) flames in a 40 kWth laboratory swirl burner

are computationally investigated under oxy-combustion conditions. A Eulerian-Eulerian approach for

modelling the two-phase flow is used. Turbulence is modelled within a URANS framework, using the

standard k-ε model and RSM.

It is observed that the RSM captures the unsteady dynamics of the coherent structures, while

latter are not captured by the k-ε model, for both flames. For the biomass flame, a longer flame is

of
predicted compared to the coal. This is expected to be mainly due to the difference in the pyrolysis rate

ro
coefficients, i.e. the larger activation energy and the smaller pre-exponential factor of biomass

-p
compared to coal, which imply a lower pyrolysis rate for the former. On the other hand, biomass has a
re
higher content on volatiles, which can burn faster compared to solid char. Of its own, this would
lP

enhance the ignition and combustion of biomass. However, in the present case, the pyrolysis rate

seems to play the dominating role, and impede the overall conversion. The analysis of the existing
na

empirical pyrolysis data and their assessment for the present type of biomass is an issue, which will be
ur

considered in the future work to increase the confidence in this respect. The predicted velocities are
Jo

compared with the available velocity measurements. For both flames, it is observed that the RSM

predictions are generally in a better agreement with the measurements compared to the k-ε model

results.

The inferior performance of the k-ε model is in line with our previous experience [33-36] in

swirling flows. On the other hand, a better performance of the k-ε model was reported in the previous

studies of other authors [18,19,21,23] on pulverized coal flames of higher power (60 kWth, 100 kWth)

in similar swirl burners. This different performance of the model may be attributed to differences in

geometries and boundary conditions.

54
Although RSM performs better than the k-ε model, its accuracy is still not very satisfactory for

the presently considered flames. For achieving a better prediction quality an analysis with an LES

framework is planned for the future, which will, however, require a considerably larger computational

effort due to higher resolution requirements.

As means of verification, the coal flame is additionally calculated using a standard/classical

Eulerian-Lagrangian two-phase formulation, using the k-ε model, leading to quite similar results to the

present Eulerian-Eulerian formulation.

of
ro
References

-p
1. M. Lackner, F. Winter, A. K. Agarwal (Eds), Handbook of Combustion (Wiley, Hoboken,2010).
re
2. A. Hussain, U. Pervez, K. Ullah, C. H. Kim, N. Asghar, Long-term scenario pathways to assess the
lP

potential of best available technologies and cost reduction of avoided carbon emissions in an

existing 100% renewable regional power system: a case study of Gilgit-Baltistan (GB), Pakistan,
na

Energy, (2021) 119855, doi.org/10.1016/j.energy.2021.119855.


ur

3. W. Suwala, Modelling adaptation of the coal industry to sustainability conditions, Energy 33


Jo

(2008) 1015-1026.

4. W. Suwala, A. Wyrwa, T. Olkuski, Trends in coal use – global, EU and Poland, IOP Conference

Series: Material Science and Engineering 268 (2017) 012003.

5. C. Shi, C. Ji, Y. Ge, S. Wang, J. Yang, H. Wang, Effects of split direct-injected hydrogen strategies

on combustion and emissions performance of a small-scale rotary engine, Energy 215 (2021)

119124.

6. S. Das, P. K. Sarkar, S. Mahapatra, Single particle combustion studies of coal/biomass fuel

mixtures, Energy 217 (2020) 119329, doi.org/10.1016/j.energy.2020.119329.

55
7. D. Iribarren, M. Martin-Gamboa, Z. Navas-Anguita, D. Garcia-Gusano. J. Dufour, Influence of

climate change externalities on the sustainability-oriented prioritization of prospective energy

scenarios, Energy, 196 (2020) 117179, doi.org/10.1016/j.energy.2020.117179.

8. P. Gladysz, W. Stanek, L. Czarnowska, G. Wecel, O, Langorgen, Thermodynamic assessment of an

integrated MILD oxyfuel combustion power plant, Energy 137 (2017) 761-764.

9. M. Qing, B. Jin, J. Ma, X. Zou, X. Wang, C. Zheng, H. Zhao, Thermodynamic and economic

performance of oxy-combustion power plants integrating chemical looping air separation, Energy

of
206 (2020) 118136, doi.org/10.1016/j.energy.2020.118136.

ro
10. P. Gladysz, W. Stanek, L. Czarnowska, S. Sladek, A. Szlek, Thermo-ecological evaluation of an

-p
intergrated MILD oxy-fuel combustion power plant with CO2 capture, utilization, and storage - A
re
case study in Poland, Energy 144 (2018) 379-392.
lP

11. T. Zhao, Z. Liu, A novel analysis of carbon capture and storage (CCS) technology adoption: An

evolutionary game model between stakeholders, Energy 189 (2019) 116352,


na

doi.org/10.1016/j.energy.2019.116352.
ur

12. W. Gajewski, A. Kijo-Kleczkowska, J. Leszczynski, Analysis of cyclic combustion of solid fuels,


Jo

Fuel 88 (2009) 221-234.

13. J. S. Leszczynski, Z. Bis, W. Gajewski, Evaluation of structure and particle velocity distribution in

circulating fluidizef beds, Powder Technology 128 (2002) 22-35.

14. B. Epple. R. Leithner, W. Linzer, H. Walter (Eds), Simulation von Kraftwerken und Feuerungen

(Springer, Vienna, 2012).

15. A. C. Benim, P. Stegelitz, B. Epple, Simulation of the two-phase flow in a laboratory coal

pulveriser, Forschung im Ingenieurwesen – Engineering Research, 69 (2005) 197-204.

56
16. J. P. Kim, U. Schnell, G. Scheffknecht, A. C. Benim, 2007, Numerical modelling of MILD

combustion for coal, Progress in Computational Fluid Dynamics – An International Journal, 7(6)

(2007) 345-361.

17. P. Madjeski, Coal Combustion Modelling in a Frontal Pulverized Coal-Fired Boiler, E3S Web of

Conferences, 46 (2018) 00010.

18. D. Toporov, P. Bocian, P. Heil, A. Kellermann, H. Stadlerm S. Tschunko, M. Förster, R. Kneer,

Detailed investigation of a pulverized fuel swirl flame in CO2/O2 atmosphere, Combustion and

of
Flame, 155 (2008) 605-618.

ro
19. L. Chen, A. F. Ghoniem, Simulation of Oxy-Coal Combustion in a 100 kWth Test Facility Using

-p
RANS and LES: A Validation Study, Energy & Fuels 26 (2012) 4783−4798.
re
20. B. M. Francetti, F. C. Marincola, S. Navarro-Martinez, A. M. Kempf, Large eddy simulation of a
lP

100 kWth swirling oxy-coal furnace, Fuel 181 (2016) 491-502.

21. P. Warzecha and A. Boguslawski, LES and RANS modeling of pulverized coal combustion in swirl
na

burner for air and oxy-combustion technologies, Energy, 66 (2014) 732-743


ur

22. A. Sadiki, S. Agrebi, M. Chrigui, A. S. Doost, R. Knappstein, F. Di Mare, J. Janicka, A.


Jo

Massmeyer, D. Zabrodiec, J. Hees, R. Kneer, Analyzing the effects of turbulence and multiphase

treatment on oxy-coal combustion process predictions using LES and RANS, Chemical

Engineering Science,166(2017)283-302

23. P. Gaikwad, H. Kulkarni, S. Sreedhara, Simplified numerical modelling of oxy-fuel combustion of

pulverized coal in a swirl burner, Applied Thermal Engineering 124 (2017)734-745.

24. A. C. Benim, B. Epple, B. Krohmer, Modelling of pulverised coal combustion by a Eulerian-

Eulerian two-phase flow formulation, Progress in Computational Fluid Dynamics – An

International Journal, 5(6) (2005), 345-361.

57
25. B. Epple, W. Fiveland, B. Krohmer, G. Richards, A. C. Benim, Assessment of-two-phase flow

models for the simulation of pulverized coal combustion, International Journal of Energy for a

Clean Environment, Vol. 6., No. 3 (2005) 267-287.

26. F. Moukalled, L. Mangani, M. Darwisch, The Finite Volume Method in Computational Fluid

Dynamics, (Springer, Berlin, 2016).

27. ANSYS Fluent Theory Guide, Release 2018 (ANSYS Inc., Canonsburg, 2018).

28. S. R. Turns, An Introduction to Combustion, 3rd ed (McGraw-Hill, New York, 2012)

of
29. M. N. Özisik, Radiative Transfer and Interactions with Conduction and Convection (Wiley,

ro
Hoboken, 1973)

-p
30. S. Bhattacharyya, H. Chattopadhyay, A. C. Benim, Computational investigation of heat transfer
re
enhancement by alzternating inclined ribs in tubular heat exchanger, Progress in Computational
lP

Fluid Dynamics, An International Journal, 17(6) (2017) 390-396.

31. S. Bhattacharyya, H. Chattopadhyay, A. Guin, A. C. Benim, Investigation of inclined turbulators


na

for heat transfer enhancement in a solar air heater, Heat Transfer Engineering 40(17-18) (2019)
ur

1451-1460.
Jo

32. B. E. Launder, D. B. Spalding, The numerical computation of turbulent flows, Comput. Meths.

Appl. Mech. Eng, 3(2) (1974) 269-289.

33. A. C. Benim, Finite element analysis of turbulent swirling flows, International Journal for

Numerical Methods in Fluids, Vol. 11, No. 6 (1989) 697-717.

34. J. L. Xia, B. L. Smith, A. C. Benim, J. Schmidli and G. Yadigaroglu, "Effect of Inlet and Outlet

Boundary Conditions on Swirling Flows" , Computers & Fluids 26(8) (1997) 811-823.

35. A. C. Benim, A. Nahavandi, URANS and LES analysis of turbulent swirling flows, Progress in

Computational Fluid Dynamics, An International Journal 5(8) (2005) 444-454.

58
36. A. C. Benim, S. Iqbal, F. Joos, A. Wiedermann, Numerical analysis of turbulent combustion in a

model swirl gas turbine combustor, Journal of Combustion, Article ID 2572035 (2016)

https://doi.org/10.1155/2016/2572035.

37. P. A. Durbin, B. A. P. Reif, Statistical Theory and Modeling for Turbulent Flows, 2nd ed. (Wiley,

Hoboken, 2011).

38. C. G. Speziale, S. Sarkar, and T. B. Gatski, Modelling the Pressure-Strain Correlation of

Turbulence: An Invariant Dynamical Systems Approach, Journal of Fluid Mechanics 227 (1991)

of
245–272.

ro
39. W. Zinser, “Zur Entwicklung mathematischer Flammenmodelle für die Verfeuerung technischer

-p
Brennstoffe”, Fortschr.-Ber. VDI Reihe 6, Nr. 171 (VDI-Verlag, Düsseldorf, 1985).
re
40. K. H. Hemsath, “Zur Berechnung der Flammenstrahlung”, Dissertation, University of Stuttgart
lP

(1969).

41. H. C. Hottel, A. F. Sarofim, The effect of gas flow patterns on radiative transfer in cylindrical
na

enclosures, International Journal of Heat and Mass Transfer, Vol. 8 (1965) 1153-1169.
ur

42. T. F. Smith, Z. F. Shen, J. N. Friedman, “Evaluation of coefficients for the weighted sum of gray
Jo

gases model”, Journal of Heat Transfer, Vol. 104 (1982) 602-608.

43. C. Yin, L. C. R. Johansen, L. A. Rosendahl, S. K. Kaer, “New weighted sum of gray gases model

applicable to computational fluid dynamics (CFD) modeling of oxy-fuel combustion: derivation,

validation, and implementation, Energy & Fuels, Vol. 24 (2010) 6275-6282.

44. C. K. Westbrook and F. L. Dryer, Simplified reaction mechanisms of the oxidation of hydrocarbon

fuels in flames, Combustion Science and Technology, Vol. 27, (1981) 31-43.

59
45. B. F. Magnussen, B. H. Hjertager, “On mathematical models of turbulent combustion with special

emphasis on soot formation and combustion”, Proc.16th Symp. (Int.) Combustion (The Combust.

Inst., Pittsburgh, PA, 1976) pp.719-729

46. S. Badzioch, P. G. W. Hawskley, Kinetics of thermal decomposition of pulverized coal particles,

Int. Eng. Chem. Proc. Des. Develop., Vol. 9, (1970), 521-530.

47. M. A. Field, F. W. Gill, B. B. Morgan, P. G. W. Hawksley, Combustion of Pulverized Coal (The

British Coal Utilization Assoc., Leatherhead, 1967).

of
48. M. M. Baum, P. J. Street, Predicting the combustion behaviour of coal particles, Combustion

ro
Science and Technology, Vol. 3 (1971) 231.243.

-p
49. D. B. Spalding, The ‘shadow’ method of particle-size calculation in a two-phase combustion,
re
Proceedings of the Nineteenth Symposium (International) on Combustion (The Combustion
lP

Institute, Pittsburgh, 1982) 941-952.

50. T. J. Barth, D. C. Jespersen, The Design and Application of Upwind Schemes on Unstructured
na

Meshes, AIAA Paper 89-0366 (1989).


ur

51. D. Zabrodiec, A. Massmeyer, J. Hees, O. Hatzfeld, R. Kneer, Flow pattern and behavior of 40 kWth
Jo

pulverized torrefied biomas flames under atmospheric and oxy-fuel conditions, Renewable and

Sustainable Energy Reviews (2020), doi:10.1016/j.rser.2020.110493.

52. D. Zabrodiec, A. Massmeyer, J. Hees, O. Hatzfeld, R. Kneer, Experimental dataset: 40 kWth

pulverized torrefied biomass and lignite flames under atmospheric and oxy-fuel conditions, Report

of RWTH Aachen, RWTK-2020-07503 (2020), doi: 10.18154/RWTH-2020-07503.

53. A. H. Lefebvre, V. G. McDonnel, Atomization and Sprays, 2nd ed, CRC Press, Boca Raton, 2017).

60
54. H. Tolvanen, L. Hokko, R. Raiko, Fast pyrolysis of coal, peat, and torrefied wood: mass loss study

with a drop-tube reactor, particle geometry analysis and kinetics modelling, Fuel 111(2013) 148-

156

55. L. D. Smoot, D. T. Pratt (Eds.), Pulverized Coal Combustion and Gasification (Plenum Press, New

York, 1979).

56. M. P. Escudier, J. J. Keller, Recirculation in swirling flow: A manifestation of vortex breakdown,

AIAA Journal, Vol. 23, No. 1 (1985) 111-116.

of
57. S. E. Gant, Reliability issues of LES-related approaches in an industrial context, Flow Turbulence

ro
and Combustion 84 (2010) 325-335.

-p
58. A. J. Banko, J. K. Eaton, A frame-invariant definition of the Q-criterion, Annual Research Briefs
re
2019, Center for Turbulence Research, Stanford University (2019) 181-194.
lP

59. M. Vanierschot , G. Ogus, Experimental investigation of the precessing vortex core in anular

swirling jet flows in the transitional regime, Experimental Thermal and Fluid Science 106 (2019)
na

148-158.
ur

60. C. T. Crowe, J. D. Schwarzkopfm M. Sommerfeld, Y. Tsuji, Multiphase Flows with Droplets and
Jo

Particles, 2nd ed (2012, CRC Press, Boca Raton).

61. A. D. Gosman and E. Ioannides, Aspects of computer simulation of liquid-fuelled combustors, J.

Energy 7(6) (1983) 482-490.

61
Highlights

 Swirling pulverized fuel flames of 40 kWth are simulated

 Biomass flame is longer than coal flame. This is partially caused by the

different pyrolysis rates

 With URANS, RSM performs better than k-ε model. Accuracy is still not

satisfactory and calls for LES

f
oo
 An equilibrium Eulerian-Eulerian two-phase formulation (EE) is proposed

r
 The EE agrees with the Eulerian-Lagrangian (EL) method and requires less

CPU time
-p
re
lP
na
ur
Jo
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

of
ro
-p
re
lP
na
ur
Jo

You might also like