Download as pdf or txt
Download as pdf or txt
You are on page 1of 35

Journal of Turbulence

ISSN: (Print) 1468-5248 (Online) Journal homepage: http://www.tandfonline.com/loi/tjot20

A comparative study of turbulence models for


non-premixed swirl-stabilized flames

Mohammad Safavi & Ehsan Amani

To cite this article: Mohammad Safavi & Ehsan Amani (2018): A comparative study of
turbulence models for non-premixed swirl-stabilized flames, Journal of Turbulence, DOI:
10.1080/14685248.2018.1527033

To link to this article: https://doi.org/10.1080/14685248.2018.1527033

Published online: 24 Sep 2018.

Submit your article to this journal

Article views: 17

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tjot20
JOURNAL OF TURBULENCE
https://doi.org/10.1080/14685248.2018.1527033

A comparative study of turbulence models for non-premixed


swirl-stabilized flames
Mohammad Safavi and Ehsan Amani

Mechanical Engineering Department, Amirkabir University of Technology, Tehran, Iran

ABSTRACT ARTICLE HISTORY


The accuracy of turbulent swirl-stabilized flame simulation strongly Received 10 May 2018
depends on the choice of turbulence model. In this study, four 3D Accepted 17 September 2018
unsteady turbulence closures, including large eddy simulation, scale- KEYWORDS
adaptive simulation, and two detached eddy simulation variants, Swirl; non-premixed;
along with four RANS models, including RNG k−ε, SST k−ω, transi- turbulent combustion; LES;
tion SST, and RSM, are examined for moderate- and high-swirl case Hybrid
studies. It is observed that the scale-adaptive simulation provides
the most accurate results for almost all variables and both swirl con-
ditions in the reactive flow. Only the 3D unsteady models predict
the vortex breakdown bubble and flame attachment state correctly.
However, based on our error analysis, the flow and composition fields
predicted by the RANS models are in acceptable agreement with the
experimental fields, especially the ones of transition SST when higher
swirl number cases or minor species concentration are of interest.
Moreover, it is concluded that the viscosity ratio criterion is a better
measure of the local LES quality than the turbulent kinetic energy
ratio, and the accuracy of a hybrid simulation may be much more
dependent on the ability of the model to operate close to the RANS
mode where the grid resolution is not sufficient for a resolving sim-
ulation than the fraction of the resolved kinetic energy. Finally, the
propriety of the base (RANS) model of a DES for the application of
interest is important, such that DES with realizable k−ε outperforms
the commonly used DES with SST k−ω model.

1. Introduction
By the development of numerical combustion models and advancement of computational
power, the design process of combustion systems has relied more on computational fluid
dynamics (CFD) simulations [1]. In order to achieve a reliable design, CFD techniques have
to precisely model the flow field and combustion process of these systems. As the swirl flow
is widely used to stabilize flames in such combustion systems, it is essential to understand
the effect of chosen models on the accuracy of the swirl-stabilized flame simulation [2].

CONTACT Ehsan Amani eamani@aut.ac.ir Mechanical Engineering Department, Amirkabir University of


Technology, 424 Hafez Avenue, P.O.Box: 15875-4413, Tehran, Iran
Supplemental data for this article can be accessed here. https://doi.org/10.1080/14685248.2018.1527033

© 2018 Informa UK Limited, trading as Taylor & Francis Group


2 M. SAFAVI AND E. AMANI

Apart from selecting an appropriate reaction mechanism, the accuracy of these simula-
tions depends primarily on the choice of turbulence closure [3] and turbulence-chemistry
interaction (or turbulent combustion) model [4].
Compared to turbulence-chemistry interaction models [5,6], fewer studies have com-
pared the effects of turbulence models on the accuracy of swirl-stabilized combustion
simulations. Reynolds-averaged Navier–Stokes (RANS), large eddy simulation (LES), and
hybrid RANS/LES methods are the common models used to calculate the turbulence of
swirling flames. Widenhorn et al. [7] indicated that the SST scale-adaptive simulation
(SAS) model performs better than SST detached eddy simulation (DES) in the prediction of
non-reactive flow in a model gas turbine combustor. West et al. [8] reported that SST-DES
model performs better than LES and unsteady RANS models in predicting the mean and
fluctuating velocity components in the simulation of non-reactive flow in swirl-stabilized
combustors. Weber et al. [9] compared the results of the Reynolds stress model (RSM),
k−ε, and an algebraic stress model (ASM) for the swirling flow in the near-burner zone.
They observed that the k−ε model brings about a considerable error in the region of burner
quarl, where the inviscid expansion of flow occurs. However, RSM and ASM models with
implementing quadratic upstream differencing (QUICK) discretization scheme for con-
vective terms are able to predict swirling vortices with adequate accuracy. Using a priori
and a posteriori tests, Hiraoka et al. [10] compared different sub-grid-scale models for LES
of premixed jet flames. Gupta and Kumar [11] studied the three-dimensional swirl flow
in a cyclone using particle tracking velocimetry and k−ε model. Their result indicates the
RNG k−ε model is more accurate in predicting the flow field compared to the standard k–
ε model.
In the case of reactive swirling flows, Navarro-Martinez and Kronenburg [12] investi-
gated the hydrogen-methane flame of Sydney burner (HM1 flame) with LES and RANS
model, coupled with conditional moment closure (CMC), based on a detailed chemistry.
They observed that the models fail to predict the scalar fields of major species and temper-
ature in the outer shear layer where the mixing between air and hot recirculating products
takes place. However, the LES-CMC results show better agreement with experimental val-
ues compared to the RANS-CMC simulation. Christo and Dally [13] studied the turbulent
non-premixed methane/hydrogen flame using different k−ε models. They observed that
the results of the standard k−ε model along with modified dissipation constant show the
best agreement with experimental values. Gran et al. [3] demonstrated that RSM is able
to give a better prediction for spreading rate of fuel jet in recirculation zone compared
to the k−ε model in the bluff-body stabilized non-premixed flame. Boudier et al. [14]
showed that the LES simulation provides a better representation of the temperature field
in gas turbines compared to RANS simulation. Wenger et al. [15] compared the results
of unsteady RANS simulation with LES simulation and experimental data. They observed
that unsteady RANS model is able to accurately model velocity of flow and resolve the
vortex core frequency. However, this model significantly underpredicts the energy con-
tained in the motion of the precessing vortex core compared to the LES counterpart. FU
et al. [16] studied the swirl-type regeneration burner with four kinds of 2-equation RANS
models. They concluded that the realizable k−ε and SST k−ω models can better simu-
late the flow field. Hu et al. [17] implemented various sub-grid models for LES simulation
of non-premixed methane flame of Sydney burner (SM1 flame). Their results show that
dynamic kinetic energy (DKE) and second-order moment (SOM) sub-grid-scale stress
JOURNAL OF TURBULENCE 3

models produce closer results to the experimental data. Chen et al. [18] compared var-
ious turbulence models (standard and RNG k−ε, SST k−ω and LES) for simulation of
reactive swirling flow in the burner quarl. Their results indicate that LES simulation out-
performs RANS models in the case of flow and scalar fields. They also concluded that SST
k−ω model gives better results among the RANS models. Sadasivuni et al. [19] compared
the result of unsteady SST with SST-SAS turbulence models combined with the scalar dissi-
pation rate (SDR) as the turbulent combustion model for simulation of the swirl-stabilized
flames. They observed that unsteady SST model shows flame elongation behavior in the
axial direction and SST-SAS simulation underpredicts the species and temperature. Luo
et al. [20] developed a new dynamic second-order moment closure for LES and evaluated
it for non-premixed swirl-stabilized flames. Wang et al. [21] showed that LES simulation
on a medium density grid is able to model the precessing vortex core structures in both
reactive and non-reactive flow of unconfined swirl flow with sufficient accuracy.
In spite of the knowledge offered by the aforementioned works, there are many questions
in selecting appropriate turbulence models which can be answered by a more compre-
hensive study comparing the LES, RANS, and hybrid RANS/LES simulations of swirl-
stabilized gas-turbine combustor flames, e.g. how much is the turbulent kinetic energy
resolution related to the accuracy? What is the influence of the base (RANS) formulation
on the performance of a hybrid model? What is the relation between the extent of the LES
region to the accuracy of a DES? etc. Moreover, several advanced turbulence models, e.g. 4-
equation transition SST, some DES variants, and SAS, have not been extensively assessed
and compared for this kind of application. In this study, we compare the results of vari-
ous turbulence models, including LES, RANS, and hybrids, on the accuracy of numerical
simulation for medium and high swirl reactive flows. Note that well-established experi-
mental databases are available for different turbulent non-premixed flame configurations
[22], including simple jet, bluff-body stabilized, and swirling flames. Among them, the
swirl-stabilized flame configuration considered here is of prime importance in gas-turbine
combustor applications and is a more challenging test case for assessing the validity of
turbulence models.

2. Test cases
The results of simulations using various turbulence models are compared to experimental
data of the Sydney burner [23,24]. In order to check the flow solver, the cold flow of the Syd-
ney burner with swirl number of 0.54 (N29S054 case) is simulated and validated against the
experimental data. To check the performance of the turbulence models for the reactive flow
in both moderate- and high-swirl number conditions, Sydney bluff-body swirl-stabilized
SM1 (moderate-swirl number) and SMA2 (high-swirl number) flames are simulated. The
schematics of SM1 and SMA2 flames are shown in Figure 1. The burner with D = 50 mm
diameter bluff size is placed in a wind tunnel which provides a co-flow secondary air stream
with mean velocity Ue = 20 m/s for both cases. SM1 flame uses CNG as fuel issuing from
a 3.6-mm-diameter nozzle with a velocity of Uj = 32.7 m/s. Swirling air is supplied from
a 5-mm-wide annular section with axial and tangential velocities of Us = 38.2 m/s and
Ws = 19.1 m/s imposing the swirl number of S = 0.5. For SMA2 flame, using premixed
methane-air (1:2 volume) as fuel, these parameters are Uj = 66.3 m/s, Us = 16.3 m/s and
4 M. SAFAVI AND E. AMANI

Figure 1. Schematic of SM1 (left) and SMA2 Sydney burners. The flow and scalar measuring cross-
sections are indicated by solid and dotted lines, respectively, and the dash-dotted line shows the
stoichiometric mixture fraction (Z st ) iso-surface (the data is extracted from present SAS simulations).

Table 1. Physical parameters of the swirling flames database [24].


Test case Uj (m/s) Us (m/s) Ws (m/s) Ue (m/s) S Rejet
N29S054 66 29.7 16 20 0.54 15700
SM1 32.7 38.2 19.1 20 0.5 7200
SMA2 66.3 16.3 25.9 20 1.6 15400

Ws = 25.9 producing the swirl number of S = 1.59. The specification of these cases is
given in Table 1 and the details of the burner can be found elsewhere [23,24].
As it can be observed in Figure 1, the SM1 flame is characterized by two main vorti-
cal structures. The one located adjacent to the bluff-body surface is generated by the flow
separation behind the bluff-body and is called lip recirculation zone (LRZ). The vortex
breakdown bubble (VBB) effect forms another swirl-induced recirculation region, called
central recirculation zone (CRZ), downstream near the centerline of the burner between
z = 76 mm and z = 96 mm [25]. The outer shear region takes collar-like shape as the shear
layer moves towards the centerline and it diverges downstream due to the existence of CRZ.
For SMA2 flame, an intensive LRZ region with the length of 100 mm is observed above the
bluff body [25]. Although the swirl number is above the critical value of 0.6, CRZ region
is not formed due to the high jet velocity at the centerline. In fact, the generated adverse
pressure gradient is not high enough to dominate the axial momentum of the fuel jet.
The experimental data for the velocity field was obtained using Laser Doppler Velocime-
try (LDV) by Masri et al. [23,26,27] at the University of Sydney. The measurements
of species concentration and temperature at different sections of the flames are taken
JOURNAL OF TURBULENCE 5

from Raman-Rayleigh-Laser-Induced-Fluorescence (LIF) provided by Sandia National


Laboratories [24].
Both SM1 and SMA2 flames are studied numerically by either LES [17,25,28–33] or
RANS [34–36] models. Stein et al. [33] showed that the LES simulation with steady-
flamelet modeling of combustion is able to capture the vortex breakdown instability in SM1
flame. El-Asrag and Menon [25] compared the flow field and recirculation zones of SM1
and SMA2 flame implementing LES. They reported that the LES simulation can capture
time-averaged velocity field with reasonable accuracy. De Meester et al. [35] implemented
2D axisymmetric RANS simulations using standard and non-linear k−ε models. They
demonstrated that their simulation method can reproduce numerical values precisely for
cold flow but the mixture fraction profiles deviate from the experimental data in the case
of reactive flow.

3. Mathematical model
The governing equations are averaged (or filtered) mass and momentum equations as

∂ρ ∂(ρuj )
+ =0 (1)
∂t ∂xj

  
∂(ρui ) ∂(ρui uj ) ∂p ∂ ∂ui ∂uj 2 ∂uk ∂
+ =− + μ + − δij − (ρτijt ) (2)
∂t ∂xj ∂xi ∂xj ∂xj ∂xi 3 ∂xk ∂xj

where ρ is the density, p the pressure, ui the velocity component in xi direction, μ the
dynamic viscosity (Sutherland’s law [37] is used to determine the variation of viscosity with
temperature), and τijt is the residual stress tensor in LES framework or the Reynolds stress
tensor, u i u j , in the case of RANS. This term is modeled based on a specific turbulence
closure. All dependent variables are (Favre) averaged quantities by default.
Here, the steady-flamelet model [38,39], based on primitive variables, accounts for
turbulence-chemistry interaction. In this model, the species mass fraction, Yi , and temper-
ature, T, are obtained from pre-computed lookup tables as functions of mixture fraction
mean, Z, variance, Z2 , and mean scalar dissipation, χst . These tables are computed and
stored for different values of these 3 variables in two steps. First, the standard steady-
flamelet equations, derived in references [38,40] and simplified in reference [41] for 1D
laminar counter-flow diffusion flames, are solved given specified χst values and the reaction
mechanism to obtain Yi and T. The accurate simplified chemical kinetics mechanisms with
low computational costs are appropriate for CFD simulation of combustion [42]. Here, a
30-species, 184-reaction CNG-air mechanism [43] based on GRI3.0 is used to model com-
bustion chemistry. In the second step, the averaged Yi and T are calculated, for given Z, Z2 ,
and χst values, assuming the β-PDF for the mixture fraction and delta-PDF forχst .
To use the lookup tables, the values of Z, Z 2 , and χst are required for each computational
cell. For this purpose, two transport equations for Z and Z 2 are solved as
  
∂ρZ ∂ ∂ λ μt ∂Z
+ (ρuj Z) = + (3)
∂t ∂xj ∂xj cp σt ∂xj
6 M. SAFAVI AND E. AMANI

  
∂ρZ 2 ∂ ∂ λ μt ∂Z 2 ∂Z ∂Z ε
(ρuj Z ) = − Cd ρ Z2
2
+ + + Cg μt (4)
∂t ∂xj ∂xj cp σt ∂xj ∂xj ∂xj k
where cp is the constant pressure specific heat capacity, λ the molecular conductivity, μt the
turbulent viscosity, and σt , Cg , and Cd are model constants which are equal to 0.85, 2.86,
and 2.0, respectively. For the LES simulation, the mixture fraction variance is modeled as
Z2 = Cz 2 |∇Z|2 [44] (instead of Equation (4)), where Cz is calculated dynamically from
the test filter and is the subgrid length scale as described by Balarac et al. [45]. χst is
calculated by [38]
Cχ εZ 2 Cχ (μ + μt )
χst = (RANS), χst = |∇Z|2 (LES) (5)
k ρσt
where k and ε are turbulent kinetic energy and its rate of dissipation, provided by the
turbulence closure, and Cχ = 2 is the model constant.
Eight models are examined for the turbulence closure. These models include four
unsteady 3D models, including a large eddy simulation (LES), two detached eddy sim-
ulation (DES) [46] variants, and the scale-adaptive simulation (SAS) [47,48]. The four
RANS-based turbulence methods are the (two-equation) re-normalization group (RNG)
k−ε [49], (two-equation) SST k−ω [50], (four-equation) transition SST (T-SST) [51,52]
and a Reynolds stress model (RSM).
The RNG k−ε model developed by Yakhot et al. [49] is based on the re-normalization
group theory [53]. In contrast to the constant turbulent Prandtl numbers used in the stan-
dard k−ε model, in the RNG k−ε model, these parameters are determined locally by
semi-analytical relations. In addition, due to the presence of an ad hoc source term in
the ε equation, the RNG model performs better than the standard k−ε in regions of high
strain and curvature [54]. The complete formulation and implementation for reactive flows
were also reported in reference [55]. The SST k−ω model proposed by Menter et al. [50]
combines the strengths of the k−ε model in free stream regions and k−ω model inside
boundary layers. In the model used here, a kinetic energy production limiter is applied
and the turbulent viscosity is defined based on the invariant of the strain-rate tensor rather
than the one of the rotation-rate tensor. The complete formulation can also be found on
NASA turbulence modeling resource (TMR) [56] (see SST-2003 model). The T-SST model
modified by Menter et al. [51,52] is based on the SST k−ω model while two additional
equations for the intermittency and transition onset criteria are solved. The complete for-
mulation can also be found on NASA-TMR [57] (see SST-2003-LM2009 model). The RSM
variant with Wilcox stress-ω model [58] for the pressure-strain redistribution closure (see
also WilcoxRSM-w2006 model on NASA-TMR [59] for the complete formulation) is a
7-equation model which solves the ω equation instead of ε in addition to the Reynolds
stresses and was recommended for swirling flows.
The SAS formulation presented by Menter and Egorov [47,48] introduces an additional
source term for the production of ω in the ω equation of the SST k−ω model. This extra
source term locally increases ω which leads to the decrease of the turbulent viscosity and
reduction of the decay of turbulent structures. Therefore, the local unsteadiness can grow in
the flow. The damping of the smallest resolved eddies is controlled by the minimum of the
von Karman length scale and a length scale proportional to the local grid size (square root
of a cell volume) [60]. Two detached eddy simulation (DES) models, one with realizable
JOURNAL OF TURBULENCE 7

Table 2. Turbulence closures used in this study.


Type Model Description
RANS k−ε 2-equation RNG k−ε
k−ω 2-equation SST k−ω
T-SST 4-equation transition SST
RSM 7-equation stress-ω RSM
Hybrid SAS SST k−ω based SAS
DES Realizable k−ε Realizable k−ε based DES
DES SST k−ω SST k−ω based DES
LES LES Germano’s dynamic SGS

k−ε model [61,62] and the other with SST k−ω approach proposed by Menter et al. [50],
are used. Similar to the SAS model, these DES models reduce the turbulent viscosity locally
but through the increase of the magnitude of the dissipation sink term in the k equation
by a factor, f DES , defined by [63]

fDES = max(1, RANS / LES ) (6)

where LES and RANS are LES and RANS length scales computed by

k3/2 k1/2
LES = CDES max , RANS = (realizable k − ε), RANS = (SST k − ω)
ε 0.09ω
(7)
where CDES = 0.61 is the DES constant and max is the maximum length of edges of a grid
cell [50]. Based on this formulation, in flow regions where RANS / LES > 1 the factor f DES
is greater than unity and the model switches to the LES mode (increased sink term) and
where RANS / LES < 1 the factor f DES equals unity and the model operates at the RANS
mode. The last unsteady 3D model is an LES model with the well-known dynamic subgrid-
scale (SGS) model proposed by Germano [64]. The local filter size is considered as the cubic
root of the cell volume. To avoid excessive oscillations of the dynamic model constant, the
local value of the numerator and denominator of the model relation are averaged over all
neighboring cells, i.e. smoothed by the test filter. Also, the value of the constant is limited
between 0 and 0.23. All the turbulence closures used here are summarized in Table 2.

4. Numerical method
The governing equations are solved based on a finite volume approach. For the discretiza-
tion of the momentum convection term in unsteady LES and hybrid simulations, the
bounded central differencing scheme is employed while in stationary RANS simulations,
QUICK scheme [65] is incorporated. For the convection terms in the scalar equations, the
second-order upwind discretization [66] is used. PRESTO scheme [67] is adopted to inter-
polate pressure on cell faces. The standard central difference scheme for the diffusion terms
and a bounded second-order implicit scheme for the discretization of transient terms in
the unsteady simulations are also applied. The discretized flow equations are solved with
PISO algorithm proposed by Issa [68].
The SAS, DES, and LES simulations are performed using an unsteady 3D approach.
The 3D domain, shown in Figure 2, contains a 0.8 m × 0.305 m × 0.305 m rectangular box.
To ensure fully-developed inlet boundary conditions, a 3.6-mm-diameter pipe for the fuel
8 M. SAFAVI AND E. AMANI

Figure 2. The 3D domain used for LES, DES and SAS modeling of Sydney burner.

inlet and a 5-mm-wide annulus for the swirling air inlet are attached to the box. The length
of these sections is 70 mm. The no-slip, adiabatic conditions are applied at all walls. The
standard wall function [69] is used for k−ε and RSM models and an enhanced blended
wall function [70] for k−ω models. The random flow generation technique proposed by
Smirnov et al. [71] is used to model the fluctuating velocity components at the inflow for
the 3D LES and hybrid simulations at all inlet boundaries, i.e. the jet pipe inlet, swirler
annulus inlet, and coflow. This inflow generator is a synthetic model based on the Fourier
harmonics which provides a divergence-free fluctuating velocity field satisfying given tur-
bulence Reynolds stress, length and time scales, and a modeled energy spectrum. Due to
the lack of data of turbulence level at the inlet, different levels of turbulence intensity in the
range of 2% to 10% at the inlet are tested for LES simulations to achieve the best agreement
with the measured axial velocity profile along the centerline [25]. The turbulent length
scale is also computed based on the hydraulic diameter. We concluded that a turbulence
intensity of 5% at the inlets results in a closer agreement with the experiment (see also the
supplemental material). The Fourier harmonics are obtained based on the inflow turbu-
lence intensity and length scale. The turbulence length scale (l) is defined as 3.8% of the
inlet hydraulic diameter (l = 0.038Dh ) and turbulence time scale is calculated based on
a turbulent intensity of 5% with the basic model proposed by Kraichnan [72]. The num-
ber of 100 Fourier harmonics is used to generate fluctuations at inlets. The details of this
model are also described by Huang et al. [73]. The convective boundary condition which
was proposed by Pauley et al. [74] is imposed at the outlet boundary. The gradient of the
mixture fraction is set to be zero at the outlet.
A structured grid with 2.7 M cells (Figure 3(a and b)) is used for the spatial discretization
of the 3D domain. The grid resolution is finer near the swirler nozzle (0.225 mm) and gets
coarser further downstream. All 3D unsteady simulations are initialized with the solution
of steady state SST k−ω turbulence model. We used a fixed time step of 5 × 10−6 second for
these simulations which results in a mean Courant number below 0.1. The 3D simulations
are first calculated for 8 × 104 time steps ( ∼ 10 flow-through times) to reach the stationary
condition and then averaged for additional 2 × 104 time steps ( ∼ 2.5 flow-through times)
to obtain time-averaged results. For the RANS models, a 2D axisymmetric domain con-
taining 81 k computational cells is utilized (see Figure 3(c)). The grid resolution near the
swirler nozzle is 0.15 mm.
JOURNAL OF TURBULENCE 9

Figure 3. (a) and (b): the structured grid, with 2.7 million cells, used for SAS, DES, and LES simulations.
Two views near the inlets, (c): the computational domain for 2D axisymmetric RANS simulations. The
embedded figure shows the grid resolution in the box I, near the inlets.

Table 3. The core hours for SMA2 simulation with different turbulence models.
DES Realizable
SAS LES k−ε DES SST k−ω RSM T-SST k−ε k−ω
4488.3 4430.0 4717.7 4487.7 3.0 2.8 2.4 2.5

Figure 4. Temperature profiles at different axial positions for 2D axisymmetric k−ε simulations on three
different grid resolutions.

The computing time (on a workstation Intel Core I7-6900 k 3.20G, 20 MB Cache, 128Gb
RAM) for each full domain 3D simulation is about 23 days and for a 2D simulation less
than half an hour. The core hour of each simulation is presented in Table 3.
To check the grid independence of RANS solutions, the simulations are carried out on
two additional 2D grids. As it is shown in Figure 4, the deviation between the results is
negligible by increasing the grid numbers from 81 k cells to 200 k cells. For the rest of paper,
the results of RANS simulations on the 81 k 2D-grid and the LES, DES, and SAS simulations
on the 2.7 M 3D-grid are discussed. To ensure that a sufficient grid resolution is used at the
bluff wall, the wall y+ is plotted along the radial direction in Figure 5 for SMA2 flame. As
one can see, for LES, y+ is smaller than 1 overall the wall surface and of the order of 1 for
the hybrid models (only approaches 1.2 in a very small region near the central jet edge).
Therefore, no wall treatment is required.
10 M. SAFAVI AND E. AMANI

Figure 5. The normalized wall distance, y+ , vs. radial direction at the bluff body wall for (a) 3D LES and
hybrid simulations, (b) 2D axisymmetric RANS simulations. R = 25 mm is the bluff-body radius.

Figure 6. Time-averaged axial (top) and swirl (bottom) velocity profiles of the cold-flow case predicted
by different models. LDV refers to the experimental data [23,24].

In order to validate the flow solver, we simulate the cold flow case of N29S054 of Sydney
swirl burner. A number of simulation results are shown in Figure 6. As it is observed, the
velocity profiles of all 3D simulations are in a good agreement with experimental data.
However, the 2D simulation with RNG k−ε model shows a deviation in axial velocity
JOURNAL OF TURBULENCE 11

profiles near the centerline at section z/D = 0.4 since this model is unable to properly
capture the central recirculation zone generated by vortex breakdown bubble instability.
At this section, the LES simulation predicts the central jet decay faster compared to the
experimental result which has also been reported by Yang and Kær [75].

5. Results and discussion


5.1. Quality of the LES and hybrid simulations
Several measures have been proposed to check the quality of an LES or hybrid (resolving)
simulation. One measure is the fraction of total kinetic energy which is exactly resolved by
LES or hybrid models. The resolved-to-total turbulent kinetic energy ratio is defined as:

kres
M= (8)
kres + kmodeled

where kres is the resolved turbulent kinetic energy contained in the instantaneous velocity
field and kmodeled is the modeled part (SGS part for LES and RANS part for hybrids), inte-
grated over the computational domain. The modeled part of the turbulent kinetic energy
for DES and SAS simulations is obtained explicitly by solving the kinetic energy transport
equation. For LES, the sub-grid (modeled) turbulent kinetic energy is estimated based on
the eddy-viscosity model [76,77]:
 2
μsgs
ksgs = (9)
ρ CDS

where CDS is the value of the sub-grid dynamic constant. The information for all simula-
tions is reported in Table 4. According to this data, all LES simulations solve for at least
80% of the total kinetic energy on the selected grid which is in accordance with Pope’s cri-
terion [71] for a sufficiently resolved LES simulation. For a hybrid model, the value of M
can show the degree the model operates in the LES mode. Among the hybrid models, DES
k−ε operates the most and SAS operates the least in the LES mode for all simulation cases.
It would be valuable to analyze the distribution of local kinetic energy ratio which is
computed by local values of kres and kmodeled substituted in Equation (8). These distribu-
tions are illustrated for LES and hybrid simulations in Figure 7. According to this measure
(Figure 7), the LES simulations are well-resolved almost everywhere. Note that for SMA2
flame, some portion of the swirler inlet annulus and boundary layer on the outer (wind
tunnel) walls far from the flame are under-resolved (low M). For the DES k−ω and SAS,
the turbulent kinetic energy in the inner and outer shear layers of the swirler flow is under-
resolved. However, the extent of the low-M region of the SAS results is much larger. For

Table 4. The turbulent kinetic energy ratio, M, for LES and hybrid models.
DES Realizable
Case LES k−ε DES SST k−ω SAS
Non-reactive 0.81 0.73 NA 0.38
SM1 0.84 0.70 0.63 0.44
SMA2 0.81 0.41 0.24 0.21
12 M. SAFAVI AND E. AMANI

Figure 7. The local resolved-to-total turbulent kinetic energy ratio contours for LES and hybrid simula-
tions of SM1 and SMA2 flames.

SMA2 flame, the DES k−ε under-resolves the kinetic energy in the lower portion of the
LRZ in addition to the swirler shear layers.
For LES, a quality assessment can be performed by examining the viscosity ratio, μt /μ.
The less this parameter is, the more the turbulence is resolved and the less the SGS diffusion
is introduced. The contours of this ratio for the LES simulations are reported in Figure 8.
The value of viscosity ratio equal to 20, which is associated with the LES index of quality
of 0.8 [78], is indicated by a dashed line in this figure for the two flames. As it can be seen
in Figure 8, μt /μ is below 20 (the index of quality is above 0.8) everywhere for the LES of
SMA2. For the LES of SM1, in most regions, μt /μ is below 20. Nevertheless, in the swirler
shear layer, after the necking region prevails, there are several spots of high viscosity ratio
(μt /μ > 20).
Note also that an extra analysis based on the power spectrum density [79] is also per-
formed in the supplementary material to check the resolution of the LES simulations. In
the hybrid RANS/LES modeling, like DES, the main part of the attached boundary layer
is resolved by RANS, while LES is applied in the regions of separated flow [80]. The other
analysis performed in this section for hybrid DES simulations is to check where the LES
and where the RANS mode is active within the computational domain. This is accom-
plished by plotting the length scale ratio, RANS / LES , defined by Equation (7), in Figure
9 ( RANS / LES > 1 shows the LES region). This figure reveals an important difference
between the two DES variants in terms of the RANS/LES mode inside the domain. The
extent of the LES region is smaller for the DES k−ε, excluding the flame far-field co-flow
JOURNAL OF TURBULENCE 13

Figure 8. The viscosity ratio contours for LES simulations of SM1 and SMA2 flames. The dashed lines
indicate the iso-level value of 20.

region in SMA2, (Figure 9) while the degree at which the DES k−ε resolves the kinetic
energy inside LES regions is higher than the DES k−ω (based on the discussion provided
through Table 4 and Figure 7). In fact, the DES k−ε switches to the RANS mode more
than the DES k−ω in the flame regions such that a large portion of the swirler shear layer
is within the RANS region for the DES k−ε simulations.

5.2. Moderate-swirl number


First, the results of the moderate-swirl case, SM1, are discussed in detail and the per-
formance of different models is evaluated and compared. Note that the acronyms ‘LES’
and ‘RSM’ used in the figures stand for the variants of LES and RSM models with Ger-
mano’s dynamic subgrid-scale [64] and Wilcox stress-ω [58] models, respectively, and
other variants of these models may perform differently. Figure 10 shows the streamlines
and axial velocity contours simulated by different turbulence models. As it can be seen
in Figure 10, all 3D unsteady simulations, i.e. LES, SAS, and DES can capture recircu-
lation zones and vortex breakdown instability well (see also Figure 1). Among the 2D
steady RANS models, only RSM model is able to predict the VBB effect and the others,
i.e. RNG k−ε, SST k−ω, and T-SST models, show similar behavior and cannot capture
the CRZ and VBB phenomena. However, RSM model is incapable of predicting the exact
shape of CRZ which is also previously reported by Shamami and Birouk [81] in the case
of a confined swirling flow. To investigate the performance of different models in more
details, the flow field profiles obtained by the models are compared with the experimental
data.
The velocity field comparisons are given in Figures 11 and 12. To facilitate the compar-
isons, the bluff-body radius, R, is chosen to non-dimensionalize the radial coordinate, r,
14 M. SAFAVI AND E. AMANI

Figure 9. The length scale ratio, LR = RANS / LES , contours for DES simulations. The dashed line indi-
cates the boundary of LES and RANS modes ( RANS / LES > 1: LES mode).

in all figures. Therefore, r/R < 1 indicates the region with a diameter less than the bluff-
body diameter. Based on Figure 11, the overall agreement of all 3D unsteady, LES and
hybrids, models (Figure 11(top)) and 2D RANS models (Figure 11(bottom)) with the
experimental data is satisfactory. However, there are some shortcomings in each of these
results. For instance, All RANS models except RSM have a poorer agreement in the region
0.2 < r/R < 0.6 at the two last axial sections where the CRZ appears. This is because they
cannot capture VBB and CRZ. On the other hand, RSM results match the experimental
data in this region, however, there is a large overprediction of the axial velocity by RSM
near the centerline (r/R < 0.2). This overprediction can be realized by examining Figure 10
and noting the unphysical persistence of the high-velocity jet core far downstream in RSM
results while this core rapidly decays after LRZ region in the other model predictions. This
discrepancy may be due to the round-jet anomaly [82], and RSM model constants need
modification for 2D axisymmetric swirl flows. The deficiency of the RSM model in the
prediction of the jet velocity deceleration of Sydney cold flow case was previously reported
by Radwan et al. [83] for both low and high swirl cases.
JOURNAL OF TURBULENCE 15

Figure 10. Streamlines and axial velocity contours for k−ε, RSM, and SAS (time-averaged) simulations.
Note that k−ω and T-SST models had the same behavior as k−ε, and LES and DES streamlines were
similar to SAS and not shown in this figure.

Figure 11. Mean axial velocity profiles of SM1 flame at different axial positions. The full 3D unsteady
simulations (top) and 2D axisymmetric RANS simulations (bottom). LDV refers to the experimental data
[23,24].
16 M. SAFAVI AND E. AMANI

Figure 12. Mean swirl velocity profiles of SM1 flame at different axial positions.

Figure 13. RMS of axial velocity (top) and swirl velocity (bottom) for the SM1 flame.
JOURNAL OF TURBULENCE 17

As it can be observed from the experimental measurements in Figure 11, the peak of
axial velocity moves from r/R ∼ 1.1 near the nozzle towards the centerline at r/R ∼ 0.8 as
the flow goes downstream, generating a bell-shaped LRZ with a necking region. The shift
in the location of peak axial velocity towards the centerline is not fully captured by the
unsteady models. This slight deviation is also observed in the LES simulation of Olbricht
et al. [84] using approximately the same grid size (4 M grid). Note that as it is observed
in Figure 11(top), SAS and LES perform better than both DES models in this regard. This
phenomenon is predicted more accurately by RANS models (Figure 11(bottom)).
Figure 12 shows that the mean swirl velocity profiles reproduced well by both full 3D
unsteady (Figure 12 (top)) and RANS (Figure 12 (bottom)) models except RSM. However,
the LES simulation shows errors in the prediction of swirl velocity in the core flow region
which was also observed by some other LES results [25,28,85] and may be due to the central
jet precession and shear layer instability. Nevertheless, Ansari’s LES simulation [86] shows
that the swirl velocity profile can be reproduced better by a finer mesh resolution. In the
RSM results, there is an overprediction of the swirl velocity near the centerline, which is
attributed to the intense swirling motion of the aforementioned (unphysical) central jet
core and deficiency in the flow prediction in the highly dynamic CRZ region of the 2D
simulation.
In Figure 13, the comparison of the root-mean-square (RMS) of velocity components
predicted by unsteady models and the experiment is illustrated. Note that, the RMS values
are the total RMS, including both the resolved and SGS contributions. The overall agree-
ment is satisfactory however the SAS predictions outweigh the others, especially in LRZ

Figure 14. Mean mixture fraction profiles of SM1 flame at different axial positions. The full 3D unsteady
simulations (top), 2D axisymmetric RANS simulations (bottom). LIF refers to the experimental data [24].
18 M. SAFAVI AND E. AMANI

Figure 15. The iso-level of stoichiometric mixture fraction (Z st = 0.054) predicted by different models
for SM1 flame. The dashed lines on the right-half of each subfigure show the measuring lines of scalar
data.

region. The LES overpredicts the stream-wise RMS near the centerline. However, the
largest error of the LES results is the underestimation of the swirl RMS at the outer shear
layer downstream the necking region (z/D > 1) which can be attributed to the insuffi-
cient resolution of the current LES in this region based on the μt /μ measure discussed in
section 5.1. This also suggests that the μt /μ criterion (Figure 8) is a better measure of the
LES result quality than the local M criterion (Figure 7). Note also that similar discrepancies
in the prediction of RMS fluctuations are also observed in the LES of SM1 by Asrag et al.
[25]. Figure 13 also shows that the DES with the realizable k−ε model has more accuracy
with respect the DES with SST k−ω.
The temperature and species mass fraction profiles are strongly affected by turbulence
model as well as turbulence-chemistry interaction. To isolate the effect of turbulence mod-
els from turbulence-chemistry interaction, which is modeled here using steady flamelet
assumption, it is instructive to investigate the mixture fraction profiles that are known to
be weakly affected by the chemistry in contrast to mass fractions and temperature.
Figure 14 reports the results of the models for the mixture fraction profiles. The over-
all agreements of 3D simulations are satisfactory and superior to the performance of the
RANS models. However, SAS has a better overall agreement with the experiment than the
others. LES simulation shows a slight underprediction in core regions and DES results are
accompanied with an overprediction, especially at the flame edges. The turbulence models
mainly fail to predict mixing features in the outer shear layer reigon [87] where entrain-
ment of the ambient air into the flame occurs. All RANS simulations underestimate the
JOURNAL OF TURBULENCE 19

Figure 16. Time-averaged temperature contours on mid-plane for different simulations of SM1 flame.

Figure 17. Instantaneous temperature contours on mid-plane for LES, DES, and SAS simulations of SM1
flame.

mixing rate of fuel jet and swirling air at the early stages (first two sections) which leads to
the underestimation of the rate of decay of the mixture fraction peak in the core flow and
the underprediction of the mixture fraction near the flame edges.
20 M. SAFAVI AND E. AMANI

Figure 18. Mean temperature profiles of SM1 flame at different axial positions.

Figure 19. Carbon monoxide mass fraction profiles of SM1 flame at different axial positions.
JOURNAL OF TURBULENCE 21

Figure 20. Stacked bar-plot for the errors of axial velocity, U, swirl velocity, W, temperature, T, mixture
fraction, Z, carbon dioxide and carbon monoxide, averaged on all plotted sections (see Equation (10)) of
SM1 flame simulations with different turbulence models.

Figure 21. The axial velocity contours and streamlines of SMA2 flame with different turbulence models.

To discuss these results further, the iso-level of the stoichiometric mixture fraction
(Zst = 0.054) is considered as the flame surface and is depicted in Figure 15. The measured
flame length of 120 mm reported by Masri et al. [24] is well captured by LES and hybrid
simulations. However, a slight overprediction is seen in the SAS and DES k−ω results and
a slight underprediction in the LES results. The underprediction of SM1 flame length by
22 M. SAFAVI AND E. AMANI

Figure 22. Mean axial velocity profiles of SMA2 flame at different axial positions. The full 3D unsteady
simulations (top), 2D axisymmetric RANS simulations (bottom). LDV refers to the experimental data
[23,24].

LES simulations was also reported by Asrag et al. [25]. Both DES models have error pre-
dicting the width of necking region (also see the profiles in Figure 14: section z/D = 1.1,
around Z = 0.054). Among the RANS models, the k−ω SST and T-SST models slightly
show the shorter flame length compared to the experiment while the RSM highly overes-
timates this length. Note that we also examined another variant of RSM using the linear
pressure-strain redistribution closure proposed by Launder, Reece, and Rodi (LRR) [88]
augmented by the wall-reflection model [89] and observed that this variant has the main
characteristics of the other RANS models used here (being unable to predict the VBB for
SM1 flame and predicting the short flame length for SMA2 flame) rather than the ones of
the stress-ω RSM (please, see the supplemental material). Therefore, the particular features
of RSM in our results originate from the modeling approach used for the pressure-strain
term.
The contours of time-averaged and instantaneous temperature for different turbulence
models are reported in Figures 16 and 17, respectively. A feature of SM1 flame is the
presence of a distinct spindle-shaped high-temperature region after the flame neck at CRZ
which is captured by all LES, RANS, and hybrid models. However, in the case of RSM this
region is extended far downstream due to the unphysical high-velocity core region. An
important difference between the two groups of simulations is that in the case of 2D RANS
simulations, there is another high-temperature region on the bluff-body at LRZ which is
unphysical and causes the temperature overprediction in this region of RANS results (see
also Figure 18). The full 3D unsteady simulations do not have this problem.
JOURNAL OF TURBULENCE 23

Figure 23. Mean swirl velocity profiles of SMA2 flame at different axial positions.

As it can be seen in Figure 18 (top), although the first group of models (3D unsteady
models) are in good agreements with the experiment inside the LRZ region, the flame
temperature is overpredicted at the edge of the flame, i.e. at the shear layer near the
edge of the LRZ. This may occur due to the errors induced in the modeling of shear
layer instability downstream of the burner nozzle due to grid size or the sensitivity of
the first group of models to the inflow turbulent intensity. Also, the use of flamelet
approach in conjunction with these unsteady models may not be accurate enough to pre-
dict the rate of chemical reaction in this region. The same problem was also reported
by Olbricht et al. [90] for the LES simulation of SM1 flame. Among the first group, the
predicted mean temperature profiles by SAS model show better agreement with the exper-
iment. In the case of 2D RANS simulations (see Figure 18 (bottom)), there is notable
overprediction of temperature near the burner nozzle (unphysical high-temperature
region). However, these models better predict the temperature profiles at the edge of the
flame. All RANS model shows almost the same behavior in the modeling of SM1 flame
except RSM.
Now the performance of the turbulence models in conjunction with the steady flamelet
model is examined through the study of mass fraction of a minor species through Figure 19.
The carbon monoxide profiles predicted by LES and SAS are in very well agreement with
the experiment, especially for the SAS while DES models overpredict CO at flame edges.
On the other hand, the RANS models, except RSM, underestimate the level of CO mass
fraction on different sections (see Figure 19 (bottom)). Among the RANS models, T-
SST induces smaller errors in minor species prediction and RSM brings about the worst
scenario.
24 M. SAFAVI AND E. AMANI

Figure 24. RMS of axial velocity (top) and swirl velocity (bottom) for SMA2 flame.

To summarize this section, an error analysis is made by averaging the deviation between
simulations and the experiment for each variable. This error is calculated by:
n
j=1 (|qexp − q|)
Er = n (10)
j=1 (|qexp |)

where q is a quantity (like axial velocity, temperature, etc.) obtained from the numeri-
cal simulation at the locations where the experimental values, qexp , have been measured.
The summation is performed over all experimental measurement locations. This error is
reported for each quantity in Figure 20. Some conclusions can be made from this figure;
SAS model possesses the smallest errors for all variables, DES model with realizable k−ε
considerably outperforms DES with SST k−ω model, and RNG k−ε is the most accu-
rate RANS model while all RANS model except RSM (with standard constants) perform
satisfactorily.

5.3. High-swirl number


SMA2 flame with highly swirling primary air is characterized by different flow regimes,
see also Figure 1. SMA2 flame length is about twice the length of SM1 flame (230 mm
vs. 120 mm visible flame length [24]) and it is almost cylindrical without necking region
observed in SM1 case.
The predicted streamlines and axial velocity contours of this flame are shown in
Figure 21. The full domain 3D simulations can capture the recirculation zone with suf-
ficient accuracy. The RANS models except RSM model underestimate the length of the
JOURNAL OF TURBULENCE 25

Figure 25. Mean mixture fraction profiles of SMA2 flame at different axial positions. The full 3D unsteady
simulations (top), 2D axisymmetric RANS simulations (bottom). LIF refers to the experimental data [24].

recirculation zone. For RSM, the recirculation zone is non-physically elongated far down-
stream. The lack of decay of the high-velocity core jet in RSM results still exists in the
high-swirl number test case.
The comparison of the axial and swirl velocity profiles for this flame are sketched in
Figures 22 and 23. As it can be observed, unsteady 3D simulations are in good agreement
with the experiments. Again, SAS model brings about the best fit with the experiment, espe-
cially for the swirl velocity. The swirl velocity profiles have two peaks before z = 50 mm.
One of them is located near the centerline and the other at the radial position close to
the swirler entrance. The first swirl velocity peak is only predicted by the SAS simulation.
For the DES with SST k−ω, both axial and swirl velocity profiles deviate from the mea-
surements near the centerline as the flow moves downstream, the same discrepancy that is
observed in 2D RSM predictions. It is also deduced from Figure 23 that the RANS models
except RSM perform satisfactorily in flow field prediction, in spite of the slight overestima-
tion of the swirl velocity at most of the sections. Note that the T-SST model shows more
accurate flow field results among the RANS models.
The SAS model also gives better prediction among all 3D simulations presented here
for both axial (see Figure 24 (top)) and swirl velocity (see Figure 24 (bottom)) RMS. There
is an underprediction of axial velocity fluctuations at shear layer near the bluff-body wall
(sections z/D = 0.2 and z/D = 0.4) for all models, which was also reported by Zhang et al.
[91] in their LES simulation. In this case, LES has the most accurate prediction with the
error of 0.27, estimated by Equation (10), and the errors of SAS, DES k−ε, and DES k−ω
26 M. SAFAVI AND E. AMANI

Figure 26. The iso-level of stoichiometric mixture fraction (Z st = 0.25) predicted by different models
for SMA2 flame. The dashed lines on the right-half of each subfigure show the measuring lines of scalar
data.

are 0.30, 0.33, and 0.48, respectively. The smaller error of the LES is consistent with the
μt /μ measure (Figure 8) again.
According to Figure 25, the models of the first group, except DES SST k−ω, are suc-
cessful to predict most features of the mixture fraction profiles in case of SMA2 flame. On
the other hand, the RANS simulations have the same issue as the ones for SM1 case (the
underestimation of the mixing process at the shear layer). This deficiency can be overcome
by changing the standard model constants, i.e. decreasing the turbulent Schmidt number.
For SMA2 flame, the iso-level of the stoichiometric mixture fraction is equal to
Zst = 0.25 and is plotted in Figure 26. The experimentally observed flame length of
230 mm is more accurately predicted by the SAS while other 3D resolving simulations
slightly underpredict the flame length. To assess the flame surface shape, the SAS simu-
lation could be considered as the reference due to its accurate prediction of the mixture
fraction profiles (Figure 25). All RANS models highly underestimate the flame height
except the RSM model.
The contours of temperature are reported in Figure 27. The flame length (230 mm [24])
is underestimated by all RANS models. In contrast to SM1, in SMA2 the flame is attached
JOURNAL OF TURBULENCE 27

Figure 27. Time-averaged temperature contours on mid-plane for different simulations of SMA2 flame.

Figure 28. Instantaneous temperature contours on mid-plane for LES, DES, and SAS simulations of
SMA2 flame.

to the bluff-body and generates a high-temperature region in LRZ. This is predicted by all
full 3D unsteady simulations here (see Figures 27 and 28) while all 2D RANS simulations
predict a detached flame incorrectly and are not able to capture this phenomenon. The
28 M. SAFAVI AND E. AMANI

Figure 29. Mean temperature profiles of SMA2 flame at different axial positions.

absence of high-temperature region in RANS results leads to the large underestimation of


the temperature inside LRZ as can be seen in Figure 29(bottom).
The comparison of temperature profiles is illustrated in Figure 29. All unsteady mod-
els, except DES with SST k−ω, predict the temperature profiles very well, notwith-
standing the slight temperature overprediction at the edges of LRZ by LES. Conversely,
the temperature is underestimated by RANS models inside LRZ near the flame edge
(see Figure 29(bottom)). However, the RANS results are acceptable except for the ones
of RSM.
To compare minor species prediction, the carbon monoxide mass fraction profiles
for SMA2 flame are depicted in Figure 30. As can be seen in this figure, the first
group slightly overpredicts CO values near the inlet section, however, the radial profiles
show fair agreement downstream of the burner. For all RANS simulations, the pre-
diction of species profiles is not accurate. However, the T-SST model performs better
than the others except near the bluff-body wall where it shows a lift-off behavior (see
Figure 26).
The average error of different turbulence models for SMA2 flame is shown in Figure 31.
Again, the SAS model has the least error among all turbulence models tested here. In
the case of the high-swirl SMA2 flame, SST k−ω and T-SST outperform the other RANS
models.
JOURNAL OF TURBULENCE 29

Figure 30. Carbon monoxide mass fraction profiles of SMA2 flame at different axial positions.

Figure 31. Stacked bar-plot for the errors of axial velocity, U, swirl velocity, W, temperature, T, mixture
fraction, Z, carbon dioxide and carbon monoxide, averaged on all plotted sections (see Equation (10)) of
SMA2 flame simulations with different turbulence models.

6. Conclusion
Various turbulence models are used to simulate medium and high swirl reacting flow. SM1
(S = 0.6) and SMA2 (S = 1.59) Sydney flames are used as the test cases. Four 3D unsteady
30 M. SAFAVI AND E. AMANI

models, including LES, SAS, and two DES variants, and four 2D axisymmetric stationary
RANS models, including RNG k−ε, SST k−ω, T-SST, and RSM, are investigated carefully.
The main conclusions of this study can be summarized as follows:

- All 3D unsteady models can capture VBB phenomenon. From the 2D RANS models,
only RSM captures this effect, however, the other results of the current RSM are not
acceptable and probably a calibration of the model constants is necessary.
- The 3D unsteady models correctly predict the state of attached flame in the high-swirl
SMA2 and detached flame in the moderate-swirl SM1 case while the 2D RANS models
do not.
- The flow field results predicted by all RANS, LES, and hybrid models are in good
agreement with the experiments.
- The viscosity ratio, μt /μ, criterion is a better measure of the propriety of the local
LES resolution than the fraction of the total turbulent kinetic energy, M, in our
results.
- Our results manifest that the SAS outperforms LES and DES models and is the best model
among all models investigated here according to the error analysis. Even though the
SAS results possessed the smallest turbulent kinetic energy resolution (M) among the
resolving simulations. Therefore, the accuracy of a hybrid model depends much more
upon the ability of the model to switch (or work close) to the RANS mode where
the resolution is not sufficient for a resolved simulation rather than the fraction of
resolved turbulent kinetic energy.
- The deviation of the 3D resolving simulations with hybrid models from the experimental
data is mainly observed in the outer shear layer region. In this region, the hybrid
models switch into the RANS mode. Therefore the ability of RANS models to predict
the mixing features in the outer shear layer plays a major role in the accuracy of the
simulation. It is observed that the SST k−ω formulation in SAS simulation, performs
better than DES simulations using Realizable k−ε or SST k−ω model.
- DES model with realizable k−ε is a far better combination than the commonly used DES
with SST k−ω model for swirl-stabilized flames. The length scale ratio, RANS / LES >
1, defines the LES region for DES simulations. The extent of the LES region is smaller
for the DES k−ε while the degree at which the DES k−ε resolves the kinetic energy
inside LES regions is higher than the DES k−ω. DES model with realizable k−ε more
switches to the base (RANS) mode in the outer shear layer where the grid resolu-
tion for a resolving simulation is below or near the threshold of μt /μ = 20 in the
LES simulation. In addition, its base model, i.e. realizable k−ε, performs well in this
region.
- RNG k−ε and T-SST RANS models perform satisfactorily, especially T-SST model if the
prediction of minor species like CO is of interest. However, their accuracy degrades,
especially for RNG k−ε, with increasing the swirl number. All RANS simulations
highly under-predict the flame length in the high swirl flow case.

Disclosure statement
No potential conflict of interest was reported by the authors.
JOURNAL OF TURBULENCE 31

ORCID
Ehsan Amani http://orcid.org/0000-0002-4125-685X

References
[1] Tyliszczak A, Geurts BJ. Controlled mixing enhancement in turbulent rectangular jets respond-
ing to periodically forced inflow conditions. J Turbul. 2015;16(8):742–771.
[2] Torkzadeh M, Bolourchifard F, Amani E. An investigation of air-swirl design criteria for gas
turbine combustors through a multi-objective CFD optimization. Fuel. 2016;186:734–749.
[3] Gran IR, Ertesvag IS, Magnussen BF. Influence of turbulence modeling on predictions of
turbulent combustion. AIAA J. 1997;35(1):106–110.
[4] Zhang H, Han C, Ye T, et al. Large eddy simulation of turbulent premixed combustion
using tabulated detailed chemistry and presumed probability density function. J Turbul.
2016;17(3):327–355.
[5] Mukhopadhyay S, van Oijen J, de Goey L. A comparative study of presumed PDFs for pre-
mixed turbulent combustion modeling based on progress variable and its variance. Fuel.
2015;159:728–740.
[6] Wu H, Ihme M. Compliance of combustion models for turbulent reacting flow simulations.
Fuel. 2016;186:853–863.
[7] Widenhorn A, Noll B, Aigner M. Numerical study of a non-reacting turbulent flow in a
gas-turbine model combustor. Proceedings of the 47th AIAA Aerospace Sciences Meeting
Including the New Horizons Forum and Aerospace Exposition; 2009 Jan 5–8; Orlando, Fl;
2009. p. 647.
[8] West JP, Groth CP, Hu JT. Assessment of Hybrid RANS/LES methods for gas-turbine
combustor-relevant turbulent flows. 22nd AIAA Computational Fluid Dynamics Conference;
2015. p. 2921.
[9] Weber R, Visser B, Boysan F. Assessment of turbulence modeling for engineering prediction
of swirling vortices in the near burner zone. Int J Heat Fluid Fl. 1990;11(3):225–235.
[10] Hiraoka K, Naka Y, Shimura M, et al. Evaluations of SGS combustion, scalar flux and stress
models in a turbulent jet premixed flame. Flow Turbul Combust. 2016;97(4):1147–1164.
[11] Gupta A, Kumar R. Three-dimensional turbulent swirling flow in a cylinder: experiments and
computations. Int J Heat Fluid Fl. 2007;28(2):249–261.
[12] Navarro-Martinez S, Kronenburg A. LES-CMC simulations of a turbulent bluff-body flame. P
Combust Inst. 2007;31(2):1721–1728.
[13] Christo F, Dally BB. Modeling turbulent reacting jets issuing into a hot and diluted coflow.
Combust Flame. 2005;142(1):117–129.
[14] Boudier G, Gicquel L, Poinsot T, et al. Comparison of LES, RANS and experiments in an
aeronautical gas turbine combustion chamber. P Combust Inst. 2007;31(2):3075–3082.
[15] Wegner B, Maltsev A, Schneider C, et al. Assessment of unsteady RANS in predicting swirl
flow instability based on LES and experiments. Int J Heat Fluid Fl. 2004;25(3):528–536.
[16] Fu J, Tang Y, Li J, et al. Four kinds of the two-equation turbulence model’s research on flow
field simulation performance of DPF’s porous media and swirl-type regeneration burner. Appl
Therm Eng. 2016;93:397–404.
[17] Hu L, Zhou L, Luo Y. Large-eddy simulation of the Sydney swirling nonpremixed flame and
validation of several subgrid-scale models. Numer Heat Tr B-Fund. 2008;53(1):39–58.
[18] Chen L, Ghoniem AF. Simulation of oxy-coal combustion in a 100 kWth test facility using
RANS and LES: a validation study. Energ Fuel. 2012;26(8):4783–4798.
[19] Sadasivuni S, Sanderson V, Bonaldo A, et al. Application of scalar dissipation rate modelling to
industrial burners in partially premixed regimes. 5th European Combustion Meeting; 2011.
[20] Luo K, Yang J, Bai Y, et al. Large eddy simulation of turbulent combustion by a dynamic second-
order moment closure model. Fuel. 2017;187:457–467.
[21] Wang P, Fröhlich J, Maas U. Impact of location and flow rate oscillation of the pilot jet on the
flow structures in swirling premixed flames. J Turbul 2010(11):N11.
32 M. SAFAVI AND E. AMANI

[22] TNF. Available from: http://www.sandia.gov/TNF/abstract.html.


[23] Al-Abdeli YM, Masri AR. Stability characteristics and flowfields of turbulent non-premixed
swirling flames. Combust Theor Model. 2003;7(4):731–766.
[24] Masri AR, Kalt PA, Barlow RS. The compositional structure of swirl-stabilised turbulent
nonpremixed flames. Combust Flame. 2004;137(1):1–37.
[25] El-Asrag H, Menon S. Large eddy simulation of bluff-body stabilized swirling non-premixed
flames. P Combust Inst. 2007;31(2):1747–1754.
[26] Al-Abdeli YM, Masri AR. Recirculation and flowfield regimes of unconfined non-reacting
swirling flows. Exp Therm Fluid Sci. 2003;27(5):655–665.
[27] Masri A, Kalt P, Al-Abdeli Y, et al. Turbulence–chemistry interactions in non-premixed
swirling flames. Combust Theor Model. 2007;11(5):653–673.
[28] James S, Zhu J, Anand M. Large eddy simulations of turbulent flames using the filtered density
function model. P Combust Inst. 2007;31(2):1737–1745.
[29] Dinesh KR, Jenkins K, Kirkpatrick M, et al. Modelling of instabilities in turbulent swirling
flames. Fuel. 2010;89(1):10–18.
[30] Yang Y, Kær SK, Yin C. Numerical study and validation of one swirling flame. Proceedings of
the European Combustion Meeting; 2011.
[31] Yang Z, Li X, Feng Z, et al. Influence of mixing model constant on local extinction effects and
temperature prediction in LES for non-premixed swirling diffusion flames. Appl Therm Eng.
2016;103:243–251.
[32] Ansari N, Goldin G, Pisciuneri P, et al. FDF simulation of swirling reacting flows on unstruc-
tured meshes. Proceedings of the 47th AIAA Aerospace Sciences Meeting Including the New
Horizons Forum and Aerospace Exposition; 2009 Jan 5–8; Orlando, FL; 2009. p. 109.
[33] Stein O, Kempf A. LES of the Sydney swirl flame series: a study of vortex breakdown in
isothermal and reacting flows. P Combust Inst. 2007;31(2):1755–1763.
[34] Rohani B, Saqr KM. Effects of hydrogen addition on the structure and pollutant emissions of
a turbulent unconfined swirling flame. Int Commun Heat Mass. 2012;39(5):681–688.
[35] De Meester R, Naud B, Maas U, et al. Hybrid RANS/PDF calculations of a swirling bluff body
flame (‘SM1’). Ichmt Digital Library Online; Begel House; 2009.
[36] Kapoor A, De A, Yadav R. Multi eulerian PDF transport modelling of turbulent swirling flame.
ASME 2012 Gas Turbine India Conference; American Society of Mechanical Engineers; 2012.
p. 377–386.
[37] Sutherland W. LII. The viscosity of gases and molecular force. London, Edinburgh, and Dublin
Philos Mag J Sci. 1893;36(223):507–531.
[38] Peters N. Turbulent combustion. Cambridge: Cambridge University Press; 2000.
[39] Poinsot T, Veynante D. Theoretical and numerical combustion. 3rd ed. Philadelphia: RT
Edwards; 2012.
[40] Pitsch H, Peters N. A consistent flamelet formulation for non-premixed combustion consider-
ing differential diffusion effects. Combust Flame. 1998;114(1):26–40.
[41] Pitsch H, Barths H, Peters N. Three-dimensional modeling of NOx and soot formation in
DI-diesel engines using detailed chemistry based on the interactive flamelet approach. SAE
Technical Paper; 1996.
[42] Ciottoli PP, Galassi RM, Lapenna PE, et al. CSP-based chemical kinetics mechanisms simpli-
fication strategy for non-premixed combustion: an application to hybrid rocket propulsion.
Combust Flame. 2017;186:83–93.
[43] Lu T, Law CK. A criterion based on computational singular perturbation for the identification
of quasi steady state species: a reduced mechanism for methane oxidation with NO chemistry.
Combust Flame. 2008;154(4):761–774.
[44] Pierce CD, Moin P. A dynamic model for subgrid-scale variance and dissipation rate of a
conserved scalar. Phys Fluids. 1998;10(12):3041–3044.
[45] Balarac G, Pitsch H, Raman V. Development of a dynamic model for the subfilter scalar
variance using the concept of optimal estimators. Phys Fluids. 2008;20(3):035114.
[46] Shur M, Spalart P, Strelets M, et al. Detached-eddy simulation of an airfoil at high angle of
attack. Eng Turbul Model Exp. 1999;4:669–678.
JOURNAL OF TURBULENCE 33

[47] Egorov Y, Menter F, Lechner R, et al. The scale-adaptive simulation method for unsteady
turbulent flow predictions. Part 2: application to complex flows. Flow Turbul Combust.
2010;85(1):139–165.
[48] Menter F, Egorov Y. The scale-adaptive simulation method for unsteady turbulent flow predic-
tions. Part 1: theory and model description. Flow Turbul Combust. 2010;85(1):113–138.
[49] Yakhot V, Orszag S, Thangam S, et al. Development of turbulence models for shear flows by a
double expansion technique. Phys Fluid A: Fluid Dynam. 1992;4(7):1510–1520.
[50] Menter F, Kuntz M, Langtry R. Ten years of industrial experience with the SST turbulence
model; 2003.
[51] Menter F, Langtry R, Völker S. Transition modelling for general purpose CFD codes. Flow
Turbul Combust. 2006;77(1):277–303.
[52] Langtry RB, Menter FR. Correlation-based transition modeling for unstructured parallelized
computational fluid dynamics codes. AIAA J. 2009;47(12):2894–2906.
[53] Yakhot V, Orszag SA. Renormalization group analysis of turbulence. I. Basic theory. J Sci
Comput. 1986;1(1):3–51.
[54] De Langhe C, Bigda J, Lodefier K, et al. One-equation RG hybrid RANS/LES computation of a
turbulent impinging jet. J Turbul. 2008;9:N16.
[55] Arjmandi H, Amani E. A numerical investigation of the entropy generation in and thermody-
namic optimization of a combustion chamber. Energy. 2015;81:706–718.
[56] TMR. SST-2003. Available from: https://turbmodels.larc.nasa.gov/sst.html.
[57] TMR. SST-2003-LM2009. Available from: https://turbmodels.larc.nasa.gov/langtrymenter_4
eqn.html.
[58] Wilcox DC. Turbulence modeling for CFD. 3rd ed. La Canada (CA): DCW industries; 2006.
[59] TMR. WilcoxRSM-w2006; Available from: https://turbmodels.larc.nasa.gov/rsm-stressomega.
html.
[60] Mehdizadeh A, Foroutan H, Vijayakumar G, et al. A new formulation of scale-adaptive simu-
lation approach to predict complex wall-bounded shear flows. J Turbul. 2014;15(10):629–649.
[61] Shih T-H, Liou WW, Shabbir A, et al. A new k-ε eddy viscosity model for high reynolds number
turbulent flows. Comput Fluids. 1995;24(3):227–238.
[62] Syed S, Hoffmann K. Detached eddy simulation of turbulent flow over a partially open cavity.
Proceeding of the 28th AIAA applied aerodynamics conference, 2010 28 June-1 July; Chicago,
USA; 2010. p. 5072.
[63] Langhe CD, Merci B, Dick E. Hybrid RANS/LES modelling with an approximate renormaliza-
tion group. I: model development. J Turbul. 2005;6:N13.
[64] Germano M, Piomelli U, Moin P, et al. A dynamic subgrid-scale eddy viscosity model. Phys
Fluid A: Fluid Dynam. 1991;3(7):1760–1765.
[65] Leonard B, Mokhtari S. ULTRA-SHARP nonoscillatory convection schemes for high-speed
steady multidimensional flow; 1990.
[66] Barth TJ, Jespersen DC. The design and application of upwind schemes on unstructured
meshes; 1989.
[67] Patankar S. Numerical heat transfer and fluid flow. Boca Raton: CRC Press; 1980.
[68] Issa RI. Solution of the implicitly discretised fluid flow equations by operator-splitting. J Com-
put Phys. 1986;62(1):40–65.
[69] Pope SB. Turbulent flows. Cambridge: IOP Publishing; 2001.
[70] Kader B. Temperature and concentration profiles in fully turbulent boundary layers. Int J Heat
Mass Transf. 1981;24(9):1541–1544.
[71] Smirnov A, Shi S, Celik I. Random flow generation technique for large eddy simulations and
particle-dynamics modeling. J Fluids Eng. 2001;123(2):359–371.
[72] Kraichnan RH. Diffusion by a random velocity field. The Physics of Fluids. 1970;13(1):
22–31.
[73] Huang S, Li Q, Wu J. A general inflow turbulence generator for large eddy simulation. J Wind
Eng Ind Aerod. 2010;98(10-11):600–617.
[74] Pauley LL, Moin P, Reynolds WC. The structure of two-dimensional separation. J Fluid Mech.
1990;220:397–411.
34 M. SAFAVI AND E. AMANI

[75] Yang Y, Kær SK. Large-eddy simulations of the non-reactive flow in the Sydney swirl burner.
Int J Heat Fluid Fl. 2012;36:47–57.
[76] Moeng C-H. A large-eddy-simulation model for the study of planetary boundary-layer turbu-
lence. J Atmos Sci. 1984;41(13):2052–2062.
[77] Sullivan PP, Horst TW, Lenschow DH, et al. Structure of subfilter-scale fluxes in the atmo-
spheric surface layer with application to large-eddy simulation modelling. J Fluid Mech.
2003;482:101–139.
[78] Celik I, Cehreli Z, Yavuz I. Index of resolution quality for large eddy simulations. J Fluids Eng.
2005;127(5):949–958.
[79] Lacaze G, Oefelein J. Development of quality assessment techniques for large eddy simulation
of propulsion and power systems in complex geometries. Livermore (CA): Sandia National
Laboratories (SNL-CA); 2015.
[80] Martelli E, Ciottoli PP, Bernardini M, et al. Detached-Eddy simulation of shock unsteadiness
in an overexpanded planar nozzle. AIAA J. 2017;55(6):2016–2028.
[81] Shamami K, Birouk M. Assessment of the performances of RANS models for simulating
swirling flows in a can-combustor. Open Aerosp Eng J. 2008;1:8–27.
[82] Pope S. An explanation of the turbulent round-jet/plane-jet anomaly. AIAA J. 1978;16(3):
279–281.
[83] Radwan A, Ibrahim KA, Hanafy A, et al. On RANS modeling of unconfined swirl flow. CFD
Lett. 2015;6(4):159–174.
[84] Olbricht C, Hahn F, Ketelheun A, et al. Strategies for presumed PDF modeling for LES with
premixed flamelet-generated manifolds. J Turbul. 2010;11:N38.
[85] Ranga Dinesh K, Jenkins KW, Kirkpatrick M, et al. Effects of swirl on intermittency character-
istics in non-premixed flames. Combust Sci Technol. 2012;184(5):629–659.
[86] Ansari N. Filtered density function for large eddy simulation of turbulent reacting flows on
unstructured meshes [Doctoral dissertation]. Pittsburgh, PA: University of Pittsburgh; 2012.
[87] Battista F, Troiani G, Picano F. Fractal scaling of turbulent premixed flame fronts: application
to LES. Int J Heat Fluid Fl. 2015;51:78–87.
[88] Launder B, Reece GJ, Rodi W. Progress in the development of a Reynolds-stress turbulence
closure. J Fluid Mech. 1975;68(03):537–566.
[89] Gibson M, Launder B. Ground effects on pressure fluctuations in the atmospheric boundary
layer. J Fluid Mech. 1978;86(3):491–511.
[90] Olbricht C, Ketelheun A, Hahn F, et al. Assessing the predictive capabilities of combustion LES
as applied to the Sydney flame series. Flow Turbul Combust. 2010;85(3–4):513–547.
[91] Zhang H, Mastorakos E. Modelling local extinction in sydney swirling non-premixed flames
with LES/CMC. P Combust Inst. 2017;36(2):1669–1676.

You might also like