Sei sulla pagina 1di 16

Fuel 186 (2016) 734749

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

An investigation of air-swirl design criteria for gas turbine combustors


through a multi-objective CFD optimization
M.M. Torkzadeh, F. Bolourchifard, E. Amani
Mechanical Engineering Dept., Amirkabir University of Technology, Tehran, Iran

h i g h l i g h t s
 A multi-objective optimization of a gas-turbine combustor is performed.
 An efficient model is proposed for the design optimization process.
 Entropy generation minimization is considered with other conventional targets.
 The Swirl number design criterion is revisited based on the new analysis.

a r t i c l e

i n f o

Article history:
Received 29 March 2016
Received in revised form 28 May 2016
Accepted 7 September 2016

Keywords:
Swirl number
Entropy generation minimization
Computational optimization
Combustion chamber
Gas turbine

a b s t r a c t
In this study, the non-premixed swirl-stabilized flame in a gas turbine combustor is simulated. For this
purpose, first, an efficient model for the simulation of combusting flow in real combustor liners is proposed which is appropriate for the computationally expensive design optimization process. Then, this
model is used to perform a multi-objective optimization of a real combustor based on the combustion
efficiency, pattern factor, pollutant (CO and NO) emission, and entropy generation minimization targets.
After investigating the effect of the Swirl number on each objective through response surfaces and sensitivity analysis, the optimal case is introduced by the decision making process. The results show that the
base case Swirl number, chosen based on a well-known correlation, is the best case if only the combustion efficiency is maximized. However, the overall optimal case which is the trade-off among the above
four objectives possesses a Swirl number 44 percent smaller than the base case.
2016 Elsevier Ltd. All rights reserved.

1. Introduction
Most of the energy produced is driven from fossil fuel resources.
One of the shortcomings of these sources of energy is pollutant
emission which is the researchers main concern because of its
effects on the global warming and the climate changes of the earth.
However, natural gas power generation facilities have been considered as more environmentally friendly power sources compared to
other hydrocarbon-fueled power plants. Gas turbines are high efficiency energy producing systems which can work with natural gas,
but still have emission such as NOx, CO2, and CO. Therefore, to alleviate this harmful effects researchers are working to reduce the
emission from gas turbines by changing the geometry or working
conditions of the gas turbine combustion chamber.
Corresponding author at: Mechanical Engineering Dept., Amirkabir University
of Technology (Tehran Polytechnic), 424 Hafez Avenue, P.O. Box: 15875-4413,
Tehran, Iran.
E-mail addresses: m.m.torkzadeh@aut.ac.ir (M.M. Torkzadeh), fbolourchi@aut.
ac.ir (F. Bolourchifard), eamani@aut.ac.ir (E. Amani).
http://dx.doi.org/10.1016/j.fuel.2016.09.022
0016-2361/ 2016 Elsevier Ltd. All rights reserved.

One of the most practical flame stabilization mechanisms in gas


turbine combustors is inlet air swirling flow and the Swirl number
can be considered as one of the most important design parameters
in these combustors. Numerous researches have been conducted
on swirl-stabilized combustion chambers. Huang and Yang [1]
and Gicquel et al. [2] reviewed many of these studies performed
by 2012. Here, a review of recent works after 2012 are provided.
These studies are summarized in Table 1.
Different stages of combustor design, e.g. detailed design,
involves the optimization of design parameters to meet several
design objectives. Conventionally, these objectives include efficiency, emission, pattern factor (PF), pressure loss, etc. The emission objective is usually accounted for by defining a parameter,
known as merit function, reflecting the amount of different pollutants such as CO, NO, etc. The pattern factor is a measure of
temperature non-uniformity at the outlet of combustors. It has
a direct effect on the thermal stresses exerted on the turbine
blades and is an important constraint in the design of
combustors.

735

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749


Table 1
Recent studies and developments on swirl-stabilized gas-turbine combustors.
Authors (year)

Method

Model (Turbulence/
Combustion)

Application

Findings

Wu et al. (2016) [9]

Computation

LES/Cold flow

Turbulent swirling flows in a


model gas turbine combustion
chamber geometry

The effect of outlet geometry contraction on the vortex


breakdown structure and the precessing vortex core in the
chamber was investigated

Jerzak and Kuznia


(2016) [10]

Experiment

Premixed combustion of natural


gas in a quartz tube model
combustor

The effects of the Swirl number and oxygen or carbon dioxide


content in fuel on the flame flashback and blow-off were studied

Li et al. (2016) [11]

Experiment

Premixed kerosene-fueled
combustor using lean staged
combustion

A new empirical NOx model was proposed for staged combustion

Ilbas et al. (2016)


[12]

Computation

Realizable k-e/Mixture
fraction-Probability
density function (PDF)

Non-premixed gas-fired
combustor with hydrogencontaining fuel

Flame temperature and NOx emission are highly affected by the


Swirl number

Shanbhogu et al.
(2016) [13]

Experiment

Premixed swirl-stabilized CH_4/


H_2-fueled combustor

Transitions in the average flame shape (or microstructure) under


acoustically coupled and uncoupled conditions were investigated

Taamallah et al.
(2016) [14]

Experiment

Methane and HydrogenPremixed-Can type

Flame transition to the outer recirculation zone (ORZ) is governed


by a Karlovitz number defined as the ratio of the ORZ spinning
frequency and extinction strain rate

Elbaz and Roberts


(2016) [15]

Experiment

Non-premixed methane swirl


flame

The effect of the quarl geometry on the flame structure and


emission characteristics was investigated

Zhang et al. (2015)


[16]

Computation

LES/Conditional moment
closure

Swirl-stabilized non-premixed
combustor

Model performance validation for non-premixed regime with


strong local extinction

Krieger et al.
(2015) [17]

Computation

Reynolds Stress Model/


Equilibrium flamelet

Qxy-fuel combustion in Can


type-model-combustor

The effect of oxy-fuel type on emission and pattern factor was


studied

Zhu et al. (2015)


[18]

Experiment
and
Computation

Realizable k-e/
Equilibrium nonpremixed flamelet

Non-premixed kerosene-fueled
swirl laboratory combustor

An enlarged recirculation zone with higher temperature can be


formed with the centerbody air injection, showing potentials in
enhancing flame stabilization and combustion efficiency

Santhosh and
Saptarshi
(2015) [19]

Experiment

Methane non-premixed swirling


flame

Primary and secondary transitions in the time-averaged flame


topology are observed as the Swirl number increases

Liu et al. [20]

Experiment
and
Computation

Standard k-e/Cold flow

Kerosene-fueled swirl-cup
combustor

Gap of nozzle shroud has a great impact on the ignition


performance and a minor effect on lean blow-out performance

Mahesh and
Mishra (2015)
[21]

Experiment
and
Computation

SST k-x/non-premixed
flamelet

Natural gas-fueled swirling


inverse jet coaxial combustor

Correlations for flame height and lean blow-out limits were


proposed as a function of the Swirl number

Bulat et al. (2014)


[22]

Computation

LES/Transported filtered
density function

Methane premixed combustion


in Can-annular (DLE) combustor
(Siemens SGT-100)

The CO and NOx emissions was studied

Khalil and Gupta


(2014) [23]

Experiment

Colorless Distributed
Combustion under both swirling
and non-swirling conditions

Increase in recirculation ratio and air injection velocity foster


distributed reaction conditions

Ghose et al. (2014)


[24]

Experiment
and
Computation

Realizable k-e/
Equilibrium presumed
PDF

Non-premixed kerosene-fueled
model combustor

The effect of air flow distribution, between primary and


secondary inlets, on flame structure, soot formation and radiative
heat transfer was studied

Zheng et al. (2013)


[25]

Computation

LES/Eddy dissipation
concept

Lean-premixed Syngas fueled


swirling combustor

The effects of hydrogen concentration, Reynolds number,


equivalence ratio and pressure were investigated

Hermeth et al.
(2013) [26]

Computation

LES/Dynamic thickened
flame

Real methane lean swirlstabilized gas turbine burner

The effects of attached and detached flame states on the dynamic


response of the flame are studied

Kim et al. (2013)


[27]

Experiment

Lean premixed swirl-stabilized


combustor

Mass flow rate of central recirculation zone is strongly related to


the combustion instabilities

Koyama and
Tachibana
(2013) [28]

Experiment

Low-swirl liquid-fueled burner

Potential of the liquid-fueled low-swirl burner technique for


industrial gas-turbine combustors is reported

Orbay (2013) [29]

Experiment
and
Computation

LES/G-equation flamelet

Simplified Volvo VT40 premixed


combustor

Heat release enhances the production of turbulence near the axis


of the combustor, but it does not alter the fundamental vortex
breakdown structure.

Steinberg et al.
(2013) [30]

Experiment

Perfectly premixed swirlstabilized flames

A distinct relationship was found between the flow structure


geometry and the pressure oscillation amplitude, with cases
having a helical vortex core resulting in higher pressure
oscillations

Ranga Dinesh et al.


(2012) [31]

Computation

LES/Cold flow

Sydney swirl burner

The effect of the Swirl number on intermittency in different


regions of a bluff-body swirling jet was examined
(continued on next page)

736

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749

Table 1 (continued)
Authors (year)

Method

Model (Turbulence/
Combustion)

Application

Findings

De and Acharya
(2012) [32]

Computation

LES/Thickened flame

Hydrogen-enriched laboratoryscale premixed combustor

Higher swirl strength and increase in level of premixedness make


the system more susceptible to upstream flame movement and
leads to combustion induced vortex breakdown

Franzelli et al.
(2012) [33]

Computation

LES/2S_CH4_BFER

Methane-air premixed
laboratory-scale combustor

The effect of premixed mixture stratification on the stability was


investigated

Wolf et al. (2012)


[34]

Computation

LES/Dynamically
thickened flame

JP10 premixed combustion in an


annular combustor

Azimuthal instability modes found in annular combustion


chambers were examined

In order to meet these targets simultaneously, a multi-objective


optimization should be applied. Several single- and multi- objective optimization studies on gas turbine combustion chambers
have been reported in the literature. Some examples are given in
Table 2 highlighting their objectives, design parameters, and salient findings.
In addition to conventional objectives, second-law related
objectives, like entropy generation minimization (EGM), has
become more and more important in CFD optimization of energy
systems [35]. Several studies also investigated the effects of different parameters on entropy generation in combustion chambers
[68], see also Table 2.
Todays energy consumption concerns have necessitated revisiting the existing designs by including the second-law efficiency
in conjunction with other conventional objectives in multiobjective optimization of combustors. In this study, an efficient
model for multi-objective optimization of real gas turbine combustors is proposed. This model includes the mathematical formulation in Section 2.1 and an equivalent simplified liner geometry
and boundary conditions in Section 4.1. Then, the model is used
to find the optimal Swirl number based on four goals, namely the
combustion efficiency, pollutant emission, pattern factor, and
EGM. This efficient procedure can be used in practical preliminary
design process.

equations for turbulent kinetic energy, k, turbulence frequency, x,


intermittency and momentum thickness Reynolds number is used
for the turbulence closure. The choice of the turbulence model is
justified by the prevalence of intermittency in swirl-stabilized
combustor flows, especially at moderate Swirl numbers where
vortex breakdown bubble appears downstream of the central
recirculation zone [31,55]. This choice is also carefully validated
in Section 4.2.
For turbulence-chemistry interaction, non-premixed flamelet
model [5658] based on primitive variables [58] is incorporated
where local (mean) mass fraction of species i, Y i , and temperature, T, are obtained from lookup tables generated as functions
of mixture fraction mean and variance, mean scalar dissipation,
and enthalpy. To generate and store the lookup tables, flamelet
equations [58] are solved for a 1-dimensional laminar counterflow diffusion flame. Here, a 34-species, 193-reaction skeletal
mechanism based on GRI3.0 is used to model combustion
chemistry. This mechanism is constructed by augmenting a
184-reaction CNG-air mechanism [59] by a 9-reaction NOx mechanism [60]. Radiation heat transfer is also accounted for by P-1
model [61].
The mixture density is calculated using the ideal gas law and
the specific heat of different species are determined as a function
of temperature from the 7-coefficent NASA polynomials [62].

2. Mathematical model

2.2. Multi-objective optimization

2.1. Governing equations

Swirl number is an important parameter in the design and operation of gas-turbine combustors. Swirl number is the ratio of the
tangential momentum to the axial momentum and is defined by

In this section, the governing equations used in this study are


introduced. All equations are given using Cartesian coordinate
notation where the summation convention over repeated indices
is implied. The governing equations include continuity and
momentum as

@
quj 0
@xj

r
@P @ sij
qu u

rxj i j @xi @xj

where uj is the jth component of the (mean) velocity vector, q the


density of the mixture, P the pressure and sij the effective stress
tensor. sij is defined as

sij



@ui @uj
2
@ul
 leff
leff

dij
3
@xj @xi
@xl

leff l lt

3
4

where l and lt are the molecular and turbulent viscosity, respectively. lt is determined by a turbulence closure model. Here, the
4-equation transition k  x SST model [54] which solves transport

R Ro
UWr2 dr
R
S Ri Ro 2
U rdr
Ri

where Ro and Ri are the outer and inner radius of the swirler, respectively, U is the axial velocity and W is the tangential velocity of inlet
air flow. In this study, we aim to carefully investigate the effects of
swirl velocity of inlet air on several important objectives in the
design process of a real gas-turbine combustion chamber and determine optimal Swirl number to satisfy all targets.
Three conventional objectives in the design of combustors are
combustion efficiency, pollutant emission, and Pattern Factor
(PF). Here, combustion efficiency is calculated by

gc

P
_ s
 inlets mh
_ FQF
m

outlet mhs

where Q F is the fuel heat of combustion, hs is the mixture sensible


_ is the mixture inlet/outlet mass flow rate into/out
enthalpy, and m
_ F is the fuel mass flow rate into the
of the combustion chamber. m
combustor.
Pattern factor is a measure of temperature non-uniformity at
the combustor outlet which is defined by

737

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749


Table 2
Sample studies on combustion chamber optimization.
Authors (year)

Design parameters

Optimization objectives

Applications

Findings

Chmielewski and
Gieras (2016)
[35]

Dilution holes area

Combustion efficiency,
NOx, and CO emissions
averaged in a range of
operations

Kerosene-fueled GTM120 miniature


combustor with
variable dilution holes

The NOx reduction by over 40% and double


reduction in CO emissions over the range of
operating points for variable geometry combustor
in comparison to the standard geometry were
noted

Arjmandi and Amani


(2015) [7]

The Swirl number, bluff size,


equivalence ratio, inlet fuel flow
rate, and the inlet air velocity

Entropy generation,
maximum temperature
and flame height

Mixed bluff-body swirl


stabilized nonpremixed combustor

Two common methods of entropy generation


calculation are compared. The two opposing
factors, namely chemical reaction and heat
transfer, have the main contribution to the total
entropy generation

Saddawi (2013) [36]

5 Geometry parameters,
equivalence ratio, reactant mass
flow rate, and injection strategy

Normalized efficiency
and fuel consumption

Premixed hydrogenair micro combustor

Introducing a micro-cooling channel to the


previous design resulted in augmented efficiency

Makhanlall et al.
(2013) [6]

Equivalence ratio

Entropy generation

Non-premixed swirlstabilized natural gas


BERL combustor

The contribution of radiation in entropy generation


rate was investigated in a combustor

Reddy et al. (2012)


[37]

Combustor configuration, exit port


diameter, and inlet velocities of air

The acoustic noise and


emissions of CO and
NOx

Kerosene-fueled
single-stage burner
with flameless/MILD
combustion

The acoustic noise were significantly reduced and


improved convective and radiative heat transfer
characteristics were achieved when the combustor
operation shifted to flameless combustion mode
from conventional combustion mode

Hajitaheri (2012)
[38]

The Swirl number and a burner


geometry parameter

Combustion efficiency

Non-premixed swirlstabilized natural gas


BERL combustor

The optimization suggested the Swirl number of


0.51 in comparison to 0.67 of the base case based
on the maximum efficiency

Saddawi (2012) [39]

14 geometry parameters,
equivalence ratio, and reactant mass
flow rate

Normalized efficiency
and fuel consumption

Premixed hydrogenair micro combustor

The optimum designs obtained with low numbers


of CFD optimization iterations exhibit the same
trend as those obtained with high numbers of CFD
iterations

Wankhede et al.
(2011) [40]

Flame stabilizer step geometry

NO emission

Premixed model
combustor

Using a steady RANS formulation, the Kriging


strategy successfully captures the underlying
response; however with unsteady RANS the
Kriging strategy fails to capture the underlying
response due to the existence of a high level of
noise

Burmberg and
Sattelmay (2011)
[41]

Swirl number

Flashback constraint

Premixed low
emission combustor

Guidelines for a proper aerodynamic design of gas


turbine combustors are given

Motsami et al.
(2010) [42]

Number and diameter of dilution


and secondary holes

Exit temperature profile


uniformity

Non-premixed liquidfueled atmospheric


combustor

The optimization leads to a more uniform


combustor exit temperature profile than with the
original one

Kulshreshtha et al.
(2010) [43]

Dilution zone holes

Liner wall temperature


and temperature
quality at the
combustor exit

Kerosene-fueled can
type combustor

Combustion chemistry modeling and thermal


boundary condition at walls are important in
prediction of temperature at inner regions of the
combustor

Laraia et al. (2010)


[44]

2 fuel spray injection and 2 swirler


geometry parameters

NOx and CO emissions

Lean Premixed
prevaporized
combustion chamber

The optimization was capable of a remarkable


pollutant emissions reduction (36 and 22% for
NOx and CO, respectively) even in presence of
conflicting design objectives

Fuligno et al. (2009)


[45]

5 geometry parameters defining the


size and position of holes

Pattern factor, NO
emissions, and pressure
loss

Lean premixed
methane-fueled
tabular combustor

An integrated approach for the design of small gas


turbine combustors which is characterized by the
integration of a 0-D code and CFD analyses with an
advanced multi-objective optimization algorithm
was implemented

Duchaine et al.
(2009) [46]

1 geometry parameter and 2


parameters indicating the fractions
of total air entering from swirler and
liner outer surface

Combustion efficiency
and pattern factor

Turbomeca reverseflow premixed ndecane-fueled


combustor

A new multi-objective optimization strategy based


on an iteratively enhanced metamodel (kriging)
coupled to a design-of-experiments was
implemented for a complex real combustor

Nirmolo et al. (2008)


[47]

The chamber diameter and number


of holes

Mixing quality

Non-premixed
methane combustion
in a cylindrical
chamber

Increasing the number of nozzles will improve the


mixing quality. For higher number of nozzles, the
optimum mixing quality is nearly independent of
normalized momentum flux ratio

Bhoi and Channiwala


(2008) [48]

Size and location of bluff body and


flammability limits

Maximum temperature
and flamlet uniformity

Methane-fully
premixed burner

The lowest NOx is achieved at Swirl number of 0.7

Iki et al. (2007) [49]

The size and position of the


combustion and cooling air inlet
holes

Combustion efficiency
and pressure loss

Non-premixed
methane micro
combustor

Considerable improvement in engine operation


could be achieved by inverting the air flow ratio
between the inner and outer liner of the combustor
(continued on next page)

738

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749

Table 2 (continued)
Authors (year)

Design parameters

Optimization objectives

Applications

Findings

Janiga and Thevenin


(2007) [50]

Fuel/air ratio in a primary and a


secondary inlet

CO emission

Laminar burner

Comparison of Genetic and Simplex optimization


algorithms

Xiong et al. (2005)


[51]

5 geometry parameters of the liner

Combustor weight,
vibration frequency,
and low-cycle fatigue
life constraints

Annular aircraft
engine combustor

An optimal configuration of the liner support


structures served as a design improvement for
future modifications were obtained

Gordon and Levy


(2005) [52]

Film cooling geometries

Liner wall temperature


and temperature
gradients

Kerosene-fueled BSE
Ltd. YT-175 engine
(annular reversal flow)
combustor

The predicted location of the liner wall hot spots


agrees well with the position of deformations and
cracks that occurred in the liner walls during test
runs of the combustor

Jyothishkumar and
Ganesan (2005)
[53]

Prediffuser geometric parameters


and dump gap ratio

Total pressure loss

Non-premixed liquidfueled annular


combustor

Cold losses are much more predominant than hot


losses and hence optimization of the geometry for
minimum cold losses is a must

T outlet;max  T outlet;ave
T outlet;ave  T inlet

PF

where T outlet;max and T outlet;ave are the maximum and average temperature at the outlet section of combustor, respectively, and T inlet is
the inlet air temperature.
To account for both CO and NOx emissions, a merit function is
usually defined by

f 

CO
CObase

1000
 

NOx
NOx

base

where CO and NOx in this equation stand for CO and NOx mass flow
rates out of the combustor outlet. Also, the subscript \base" refers
to the values corresponding to the base case design which is
described in Section 3.
Here, we consider another objective, i.e. entropy generation
minimization, to account for the quality of energy conversion
and reduce the energy loss. For this purpose, the total entropy generation rate in the combustor is calculated by

I
Sgen

q
Aoutlet

walls

!
I
X
Y i si uj dAj 
i

dQ
T

Ainlets

!
X
Y i si uj dAj

3. Specification of the real combustor


The geometry of the liner of the real combustor studied here is
shown in Fig. 1. Fuel, natural gas with the composition reported in
Table 3, enters the combustor from the inner annulus indicated by
word Fuel in Fig. 1 and the swirling air flow enters through the
outer annulus. The rest of air is introduced through the other holes.
The specification of the chamber is given in Table 4.
In the process of preliminary design of gas-turbine combustors,
the inlet Swirl number is chosen based on several criteria. One of
this criteria is described by the following equation [65]

LRZ  2Ro S

10

Which describes that the length of the recirculation zone, LRZ , is


affected by the Swirl number (see also Fig. 2). Based on Eq. (10),
assuming a combustor sizing, a Swirl number criterion can be proposed such that the recirculating region fills the combustor primary
zone completely [65]. For the combustor studied in this work
(Fig. 1), this Swirl number equals 1.87.
4. Numerical method

where si is the specific entropy of species i. The terms on the


right-hand side of Eq. (9) account for entropy outflow, entropy
inflow and entropy transfer from the combustor walls, respectively. Note that, the heat flux to the walls, dQ, is the sum of convective and radiative fluxes calculated by the model described in
the previous subsection.
Eqs. (6)(9) describe the 4 objectives used in our multiobjective optimization study. The optimization is done by the
Response Surface (RS) method where the objective parameters
are constructed as a function of design variables, here the Swirl
number, to fit a given set of data which is obtained by CFD solution
of the model described in Section 2.1.
Different kinds of response surfaces are used in various
applications. For noisy responses like the ones encountered in
combustion problems, Non-Parametric Regression (NPR) method
[63] would be a wise choice and results in a more satisfactory
goodness of fit as will be seen in results and discussion section.
After constructing the response surfaces, the multi-objective
optimization is performed by Non-dominated Sorting Genetic
Algorithm II (NSGA II) [64].

For the numerical discretization of the governing equations,


finite volume method is applied. The diffusion terms are discretized by the second-order central scheme and the convection
terms by the upwind scheme. The SIMPLE algorithm [66] is used
for the pressure-velocity decoupling of the Navier-Stokes equations. Due to the complexity of the problem, a small underrelaxation factor of 0.3 is required for the momentum equation.
In the next subsections, the model combustor geometry, boundary
conditions and computational grid are proposed. Then, the grid
sensitivity of the solution and the validation of the model are
investigated.
4.1. Model combustor
The problem of numerical simulation of real gas-turbine combustors involves both complex physics and complex geometries.
Therefore, 3D CFD simulation of a real combustor would be a computationally prohibitive task especially for optimization studies
where a large number of cases have to be solved. In this section,
a simplification procedure is proposed and described in detail to
obtain an axisymmetric 2D (with swirl velocity) variant of a real
combustor.
The 2D axisymmetric model combustor is shown in Fig. 3. This
efficient 2D model assumes that each row of holes is equivalent to

739

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749

Fig. 1. The geometry of the liner of the combustor studied in this work (top), sectional view of the liner (bottom). All dimensions are in millimeters.

Table 3
Natural gas composition (mole fractions).
Species

Fuel

CH4
C2H6
CO2
N2

0.9
0.085
0.005
0.01

Table 4
The specification of the combustion chamber. FOA indicates the
fraction of the total air entering the combustor from each section.

a round groove around the liner whose inlet area equals the area of
the original holes (see Fig. 4), i.e.

npD2hole =4 pDliner wr

11

where wr is the width of the groove, n the number of holes in the


row, Dhole the hole diameter and Dliner the liner diameter at the specified section. As far as the jet velocity issuing from combustor holes
has the prime effect on the dynamics of the mixing process, groove
width is selected according to Eq. (11) to achieve exactly the same
bulk velocity for 2D groves as the one for original 3D holes. In other
words, with this formulation, the groove width is smaller than the
hole diameter but results in the same inlet area and bulk velocity
as the original problem. As far as only the flow in the liner is solved,
especial care should be taken to determine boundary conditions for
the model combustor to operate at a condition as close as possible
to the one of the real combustor.
The inlet flow boundary conditions can be specified knowing
the flow direction, mass flow rate, temperature and composition
(from Table 4) at each inlet section; gaseous fuel is introduced axially, normal to inlet boundary, while the swirling air has also a tangential velocity component which is determined based on the

Design parameter

Value

Liner diameter
Fuel flow rate
Inlet fuel temperature
Number of primary holes
Diameter of primary holes
Number of secondary holes
Diameter of secondary holes
Number of dilution holes
Diameter of dilution holes
Total inlet air flow rate
Primary holes FOA
Secondary holes FOA
Dilution holes, 1st row, FOA
Dilution holes, 2nd row, FOA
Swirler FOA
Cooling air, 1st slot, FOA
Cooling air, 2nd slot, FOA
Cooling air, 3rd slot, FOA
Cooling air, 4th slot, FOA
Inlet air temperature
Inlet air pressure
Inlet air Swirl number (swirler)

245 mm
0.0712 kg/s
430 K
20
18 mm
20
14 mm
8 (two rows)
34 mm
4.288 kg/s
20%
15%
15%
15%
12%
8%
5%
5%
5%
461 K
3.899 bar
1.87

given Swirl number. The inlet flow through cooling slots is also
normal to the boundary face, parallel to the liner walls. This flow
forms an air layer on the inner liner wall to protect it from hot
combustion gaseous. It is known that the inflow from primary, secondary, and dilution holes is not normal to the hole surface. Here,

740

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749

casing walls and the equivalent thermal circuit model which is


implemented as the thermal boundary condition.
Note that Dan is the hydraulic diameter of the annulus and is
equal to Dcasing  Dliner . Thus, the conduction resistance of the wall
is

Rcond

ttbc tliner

ktbc kliner

17

where t and k are the thickness and thermal conductivity, respectively. The thermal barrier coating (TBC) protects combustor liner
from overheating and reduces the heat loss across the walls. The
radiation resistance is calculated by

Rrad

an empirical correlation [67] is used to determine the direction of


inflow from the holes as
2

CD
C D1

13

where h is the angle between the inflow and main flow direction
and C D the discharge coefficient. An estimation of the discharge
coefficient for round sharp-edged holes is provided by [68]

1:25K  1
4K  K2  b2 
2

Dan
Nu kf

19

Nu 0:05ReD;F 0:8 Prf 0:4

K!1

CD

Rconv

12

C D1 lim C D

0:5

 0:8
T an
Tf

b
b0

Tf

T an T w2
;
2

ReD;F

_ an
m
qan Aan

Dan

tf

14

15

16

where b0 is the ratio of the hole area to the annulus cross-sectional


area.
The no-slip boundary condition is applied at liner walls. To
determine the thermal boundary conditions at the liner walls,
the conduction heat across the wall, convection from the liner to
the annulus air flow and radiation from the liner to the casing
are accounted for Fig. 5 shows the configuration of the liner and

20

where Tf and Reynolds number are expressed by

where b is the ratio of the hole mass flow rate to annulus, between
liner and casing, mass flow rate, and K is the ratio of the jet dynamic
pressure to the annulus dynamic pressure upstream of the holes.
The coefficient K is approximated by the correlation [68]


q
K 1 0:64 2w2 4w4 1:56w2 4b  b2

18

where  is the emissivity coefficient and r the Stefan-Boltzmann


constant. Convective resistance is also calculated using an empirical
correlation [69] as follows

Fig. 2. The recirculating zone inside the combustor liner.

sin h

casing liner 1  casing


rliner casing T casing T w2 T 2casing T 2w2

Fig. 4. The approximation of a row of holes with a round groove.

Fig. 3. The 2D axisymmetric geometry of the model combustor liner (figure has been plotted with the scale shown at the bottom).

21

741

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749

Fig. 5. Left: the configuration of the liner wall (e=Dan 0:0275). TBC stands for thermal protection of the liner. Right: the thermal circuit model for the wall boundary
condition.

_ an and T an are local air mass flow rate and bulk air temperature
m
through the annulus and tf and Pr f are dynamic viscosity and
Prandtl number of air evaluated at T f .

used in this study, computational grid should be devised carefully


to be fine enough in high gradient regions, such as inlet areas, to
reduce grid sensitivity and coarser in other regions to save computational costs while it changes smoothly from fine to coarse
regions. This makes the simulation of large number of cases possible for optimization purposes. We examined many cases to find an
appropriate grid configuration which results in grid independent
solutions for different operating conditions of our interest. This
grid is reported in Fig. 11. This 2D grid is composed of 9159 cells
of rectangular shape in inlet annulus and near wall regions (boundary layers) and of triangular shape in inner regions. To show the
grid sensitivity, in Fig. 12, the numerical solution of the base design
(Table 4) using this grid is compared with the results using a
refined grid, with 39,714 cells. Mean deviation of temperature profile between the results of the two grids is 3.9%. The results of this
section show the level of accuracy and reliability of our
computations.

4.2. Validation and grid independence check


For validation of the model, we use Sydney swirl flame database
[70], SM1 case, which is a reliable test case for swirling flow and
combustion in gas turbine combustion chambers. The details of
the burner configuration and operating conditions have been
reported in other studies [7175]. A 2D grid with 86,000 quadratic
cells is used for the simulation of the model described in Section 2.
The comparisons of computed and experimental results for temperature, axial velocity, swirl velocity, and CO2 and CO mass fraction profiles at different distances from the inlet nozzle are
reported in Figs. 610.
As it is observed in these figures, the computed results are in
good agreement with the measurements. While many turbulence
models have problem predicting the swirling flow when a stationary approach and 2D assumption is incorporated, this validation
study shows that the present formulation including the 4equation transition k  x SST model [54] brings about satisfactory
results. This point has also been investigated in detail in our other
work [75].
Based on our experience, the swirling flow solution is highly
sensitive to grid resolution. For complex geometries, like the one

5. Result and discussion


The results of the present numerical simulation of the base case
(see Table 4) which is a high Swirl number case are discussed first.
To examine the flow field, the contours of swirl and axial velocity
components and turbulent kinetic energy are shown in Figs. 1315.
The region of high swirling flow inside the combustor is indicated
in Fig. 13. This region visualizes a large swirling flow structure

(a)

(b)

(c)

(d)

(e)

Fig. 6. Temperature versus radial direction in different cross sections with different distances from the inlet nozzle (a) x = 6.8 mm, (b) x = 20 mm (c) x = 40 mm (d) x = 70 mm
(e) x = 125 mm. Comparison between experimental measurements [70] (symbols) and present computations (lines).

742

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749

(b)

(a)

(c)

(d)

(e)

Fig. 7. Axial velocity versus radial direction in different cross sections with different distances from the inlet nozzle (a) x = 6.8 mm, (b) x = 20 mm (c) x = 40 mm (d) x = 70 mm
(e) x = 125 mm. Comparison between experimental measurements [70] (symbols) and present computations (lines).

(a)

(b)

(c)

(d)

(e)

Fig. 8. Swirl velocity versus radial direction in different cross sections with different distances from the inlet nozzle (a) x = 6.8 mm, (b) x = 20 mm (c) x = 40 mm (d) x = 70 mm
(e) x = 125 mm. Comparison between experimental measurements [70] (symbols) and present computations (lines).

attached to the swirler of the combustor. Inside this structure,


there is a region of negative axial velocity (see Fig. 14). This region
is called internal recirculating zone (IRZ) which is the characteristic
of high Swirl number cases. To further illustrate this region,
streamlines are plotted in Fig. 16. As it is observed, the IRZ region
in this case is composed of two pairs of counter-rotating cells (one
cell of each pair is shown due to the symmetry). At the interface of
IRZ and the high-swirl region, a high-shear region exists which
produces considerable values of turbulent kinetic energy as can
be seen in Fig. 15. Due to the relatively small angle of liner dome,
the expansion of inflow occurs gradually and there is no corner
recirculation zone (CRZ) inside the chamber (see Fig. 16).
Fig. 17 shows contour of the temperature in the combustor. As
it can be observed in this figure, the flame surface approaches to
the dome walls. This is because the high Swirl number causes
the high shear region to moves towards the dome walls. As men-

tioned above, this region is also the region of high turbulent kinetic
energy which promotes mixing and reaction rates (in nonpremixed regime). Therefore, the flame surface is located in this
region, near walls, and special care must be taken to prevent the
dome walls from burning. This is the reason why larger cooling
flow rates has been assigned to the dome cooling slot (see Table 4).
Figs. 18 and 19 illustrate pollutant emissions, including NO and
CO. It can be observed in these figures that CO forms first but oxidizes to CO2 due to the increase of temperature beyond 1700 K. In
this case, due to the relatively large length of the combustor the
level of CO is low at the combustor exit. However, this is not the
case for NOx emission (see Fig. 19). NOx formation is considerable
in this case due to the high temperatures. Maximum NOx exists in
the middle of combustor near the centerline, but further downstream the level of NOx decreases a bit due to the effect of air
entering from dilution holes.

743

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749

(c)

(d)

(e)

Fig. 9. CO2 mass fraction versus radial direction in different cross sections with different distances from the inlet nozzle (a) x = 6.8 mm, (b) x = 20 mm (c) x = 40 mm (d)
x = 70 mm (e) x = 125 mm. Comparison between experimental measurements [70] (symbols) and present computations (lines).

(a)

(b)

(c)

(d)

(e)

Fig. 10. CO mass fraction versus radial direction in different cross sections with different distances from the inlet nozzle (a) x = 6.8 mm, (b) x = 20 mm (c) x = 40 mm (d)
x = 70 mm (e) x = 125 mm. Comparison between experimental measurements [70] (symbols) and present computations (lines).

Fig. 11. The computational grid used in this study. Top: full view. Bottom left: near inlet. Bottom right: near dilution holes.

744

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749

(d)

(c)

(b)

(a)

Fig. 12. Comparing temperature profiles using main (solid line) and fine (dashed line) grid in different cross sections with different distances from the inlet (a) x = 0.09 m, (b)
x = 0.11 m (c) x = 0.145 m (d) x = 0.18 m.

W (m/s):

30

55

80

105

130

155

180

205

Fig. 13. Contour of the swirl velocity for the base case (S = 1.87). The contour level W = 38 (m/s) emphasized by the black line surrounds the high swirl region.

Fig. 14. Contour of the axial velocity for the base case (S = 1.87).

Fig. 15. Contour of the turbulent kinetic energy for the base case (S = 1.87).

Fig. 16. The streamlines for the base case (S = 1.87).

745

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749

Fig. 17. Contour of the temperature for the base case (S = 1.87).

Fig. 18. Contour of the mass fraction of CO for the base case (S = 1.87).

Fig. 19. Contour of the mass fraction of NO for the base case (S = 1.87).

7800

2500

7600
2300

Temperature (K)

Merit funcon

7400
7200
7000
6800
6600
6400

DOE points
Response Curve

6200

0.5

1900

Tmax outlet
Tmax chamber

1700

6000
0

2100

1.5

2.5

1500
0

0.5

Swirl number

1.5

2.5

Swirl number

Fig. 20. Left: Response Surface (RS) of the merit function, f, as a function of the Swirl number. Right: Maximum temperature of outlet and chamber versus Swirl number.

1.6
1.5
1.4

Paern factor

Now, the effect of the Swirl number on flow and combustion


characteristics is examined and the optimal condition is proposed.
The design of experiment (DOE) constitutes 22 design points to
cover low- to high-swirl conditions with the Swirl numbers in
the range 0.22.2. The numerical simulation of these design points
are performed for the model combustor proposed here and the
results are used to generate the Response Surfaces (RS) by NPR
method. The RCs for the four objectives, described in Section 2.2,
are shown in Figs. 2023.
As we can see in Fig. 20, there is a peak in merit function (minimum emissions) at S = 0.76 and above this value, merit function
reduces (emission increases) monotonically by the increase of
the swirl velocity. This trend is due to the increase of NOx by the
rise of the maximum combustor temperature when the swirl
velocity grows above the value of 0.76 as depicted in Fig. 20 (right).
According to Fig. 21, by increasing the Swirl number the pattern

1.3
1.2
1.1
1

DOE points
Response Curve

0.9
0.8
0

0.5

1.5

2.5

Swirl number
Fig. 21. RS of the Pattern Factor, PF, as a function of the Swirl number.

746

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749

Combuson eciency

0.99
0.98
0.97
0.96
0.95

DOE points
Response Curve

0.94
0

0.5

1.5

2.5

Swirl number
Fig. 22. RS of the combustion efficiency, gc , as a function of the Swirl number.

4900

DOE points
Response Curve

Sgen (W/K)

4800

4700

4600

4500

4400
0

0.5

1.5

2.5

Swirl number
Fig. 23. RS of the total entropy generation rate, Sgen , as a function of the Swirl
number.

0.1
Eciency
Sgen

Normalized eect factor

0
-0.1

Merit
funcon
Paern
factor

-0.2
-0.3
-0.4
-0.5

Fig. 24. Local sensitivity of the output parameters (targets).

factor will decrease which indicates the better mixing and more
uniform outlet temperature profile induced by the larger Swirl
numbers.

The variation of the combustion efficiency with the change of


the Swirl number follows a more complex pattern. As it is illustrated in Fig. 22, the design points suggest a noisy variation of
the combustion efficiency by the change of swirl velocity. These
noisy values are due to the small, but important, changes in combustion efficiency. All points lie in the range 0.9620.998 with
about 3.5 percent difference in efficiency and generally predict a
slight increase of this objective with the raise in the swirl velocity,
except for very low or very high Swirl numbers. As it can be
observed in this figure, the generated RS by RNN method well predicts this trend and indicates the propriety of RNN for combustion
problems.
The entropy generation RS is illustrated in Fig. 23. The RS has an
absolute minimum at S = 0.77 and there is a sharp increase in
entropy generation between S = 1.2 and S = 2.
In the following, the local sensitivity of each output parameter
is investigated to find and compare the extent that each objective
is affected by the Swirl number. The sensitivity chart is reported in
Fig. 24. The normalized effect factor is a measure of the linear association between an objective and a design variable (here Swirl
number). This degree of relationship is a number between 1
and 1. The correlation value of 1, 0 and 1 correspond to perfectly
positively correlated, uncorrelated and perfectly negatively correlated variables.
According to Fig. 24, the pattern factor is influenced most by
the change in the Swirl number and emission (merit function)
is the second most sensitive output parameter while the combustion efficiency shows the least level of sensitivity to the swirl
velocity.
To propose an optimal Swirl number, a multi-objective optimization is performed here based on the generated response surfaces. For this purpose the Pareto fronts are generated using
NSGA II algorithm through maximizing gc and f, and minimizing
PF and Sgen . Then, the best candidates are found by the process of
decision making. This is just like converting a multi-objective problem to a single-objective one by allocating a weight to every objective. In this study, Genetic Algorithm decision making method has
been utilized considering equal weights for all objectives. The three
best choices obtained from the multi-objective optimization are
listed in Table 5.
The best choices are in the range of S = 0.82 to S = 1.05. Candidate 1 with S = 1.053 is the optimal case based on the trade-off
between all objectives. This design possesses the best efficiency
and pattern factor among the three candidates. Therefore, this
multi-objective optimization proposes a Swirl number which is
44 percent smaller than the base case value obtained based on
the criterion described in Section 3. If emission is of higher importance, Candidate 3 with S = 0.821 can also be a proper choice provided that the larger pattern factor is tolerable.
It is also instructive to examine the conditions which best satisfy only one of objectives (single-objective optimization). These
conditions are listed in Table 6. According to this table, minimum
value of all objectives occurs at Swirl numbers higher than 0.75
which is a high Swirl number.
Comparing Candidate 1 of multi-objective optimization and
Table 6, it is seen that the entropy generation of the optimal case
(Candidate 1) is only 0.2% greater than the minimum entropy generation. It means that optimal case causes an entropy generation

Table 5
Candidate points obtained from present multi-objective optimization.
Parameter

Swirl number

Combustion efficiency

Merit function

Pattern factor

Entropy generation (W/K)

Candidate point 1
Candidate point 2
Candidate point 3

1.053
0.937
0.821

0.9731
0.9682
0.9638

7229.09
7437.93
7568.98

1.335
1.384
1.421

4507.80
4501.31
4498.94

747

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749

Table 6
The results of the single-objective optimization.
(Single) objective

Swirl number

Combustion efficiency

Merit function

Pattern factor

Entropy generation (W/K)

Pattern factor
Combustion efficiency
Merit function
Entropy generation

2.200
1.841
0.757
0.766

0.9896
0.9980
0.9624
0.9625

6373.17
6441.07
7589.63
7589.14

0.959
1.057
1.434
1.433

4823.74
4836.17
4498.65
4498.64

value close to the best possible entropy generation in the problem.


It is also interesting to note that the maximum combustion
efficiency corresponds to S = 1.841 which is very close the base

case value (S = 1.87). This indicates that the base case (original criterion for the Swirl number [65]) can be acceptable when only the
first-law efficiency is considered.

Fig. 25. Temperature contours of the case listed in Table 6. From top to bottom: minimum PF (S = 2.20), maximum gc (S = 1.841), multi-objective optimal (S = 1.053) and
maximum f (S = 0.757).

Fig. 26. Contour of CO (up) and NO (down) mass at the condition of maximum merit function (S = 0.757). For direct comparison of the colors of this figure with those of
Figs. 18 and 19, exactly the same legends are used for all these figures.

748

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749

The contour of the temperature for each of the cases listed in


Table 6 and the Candidate 1 from multi-objective optimization is
shown in Fig. 25. At S = 2.20 corresponding to the smallest pattern
factor the high temperature region is located well upstream of the
outlet and the mixing of hot gas and dilution stream performs over
a longer distance before the gas leaves the combustor. This promotes uniformity of the flow at the outlet. At S = 1.841 corresponding to the largest efficiency the high temperature region extends
over larger area of the combustor which indicates more intense
and complete combustion. The multi-objective optimal condition
corresponding to S = 1.053 occurs in the middle of this range of
Swirl numbers. The minimum emission corresponding to
S = 0.757 involves a slower combustion and longer reaction region
relative to the other cases mentioned in Fig. 25. This slower combustion results in decreased CO oxidation, temperature rise, and
NOx formation such that the level of NOx and CO are more balanced at the outlet (compare Fig. 26 with Figs. 18 and 19) and a larger merit function is obtained in this case. Although, this
optimization results in a lower NO levels at combustor outlet compared with base case (see Fig. 19), further improvements can be
achieved adopting this multi-objective optimization including several additional design parameters like dilution hole diameter and
flow rate.
Another interesting conclusion that can be inferred from Table 6
is that the Swirl number corresponding to the minimum entropy
generation is close to the one of minimum emission however this
condition does not coincide with and is far from the one of maximum efficiency and minimum pattern factor.
6. Conclusion
In this work, the numerical simulation of swirl-stabilized combustion in gas turbine combustors was studied by proposing an
efficient model appropriate for the design optimization process
involving a large number of simulations. For this purpose, the 4equation transition k  x SST model for turbulence closure was
used in conjunction with the steady-flamelet model for
turbulence-chemistry interaction and a proper choice for the reaction mechanism. The other part of the modeling was the approach
used for simplifying the complex geometry of a real combustor
such that only combustor liner is simulated with equivalent hydrodynamic and thermal boundary conditions detailed in Section 4.1.
At the next stage, the model was used to investigate the effect of
air Swirl number on different targets in a multi-objective optimization process. This optimization accounts for three conventional
objectives, i.e. combustion efficiency, emission, and pattern factor,
and a newer one, i.e. entropy generation minimization. Finally, the
design criteria for choosing the optimal Swirl number was discussed in detail. Since, this carefully devised 2D axisymmetric
model offers a tractable alternative to computationally expensive
3D simulations for multi-objective design problems, it would be
valuable to assess the accuracy of this model through comparison
with the results of 3D simulations in future studies.
References
[1] Huang Y, Yang V. Dynamics and stability of lean-premixed swirl-stabilized
combustion. Prog Energy Combust Sci 2009;35(4):293364.
[2] Gicquel LY, Staffelbach G, Poinsot T. Large eddy simulations of gaseous flames
in gas turbine combustion chambers. Prog Energy Combust Sci 2012;38
(6):782817.
[3] Bejan A. Entropy generation minimization: the new thermodynamics of finitesize devices and finite-time processes. J Appl Phys 1996;79(3):1191218.
[4] Amani E, Nobari M. A numerical study of entropy generation in the entrance
region of curved pipes. Heat Trans Eng 2010;31(14):120312.
[5] Amani E, Nobari M. A numerical investigation of entropy generation in the
entrance region of curved pipes at constant wall temperature. Energy 2011;36
(8):490918.

[6] Makhanlall D, Munda JL, Jiang P. Radiation energy devaluation in diffusion


combusting flows of natural gas. Energy 2013;61:65763.
[7] Arjmandi H, Amani E. A numerical investigation of the entropy generation in
and thermodynamic optimization of a combustion chamber. Energy
2015;81:70618.
[8] Jafari M et al. Inclusion of entropy generation minimization in multi-objective
CFD optimization of diesel engines. Energy 2016;114:52641.
[9] Wu Y et al. Effect of geometrical contraction on vortex breakdown of swirling
turbulent flow in a model combustor. Fuel 2016;170:21025.
[10] Jerzak W, Kuznia M. Experimental study of impact of swirl number as well as
oxygen and carbon dioxide content in natural gas combustion air on flame
flashback and blow-off. J Nat Gas Sci Eng 2016.
[11] Li L et al. Emission characteristics of a model combustor for aero gas turbine
application. Exp Therm Fluid Sci 2016;72:23548.
_ M, Karyeyen S, Yilmaz I.
_ Effect of swirl number on combustion
[12] Ilbas
characteristics of hydrogen-containing fuels in a combustor. Int J Hydrogen
Energy 2016.
[13] Shanbhogue S et al. Flame macrostructures, combustion instability and
extinction strain scaling in swirl-stabilized premixed CH4/H2 combustion.
Combust Flame 2016;163:494507.
[14] Taamallah S, Shanbhogue SJ, Ghoniem AF. Turbulent flame stabilization modes
in premixed swirl combustion: physical mechanism and Karlovitz numberbased criterion. Combust Flame 2016.
[15] Elbaz A, Roberts W. Investigation of the effects of quarl and initial conditions
on swirling non-premixed methane flames: flow field, temperature, and
species distributions. Fuel 2016;169:12034.
[16] Zhang H et al. Large eddy simulation/conditional moment closure modeling of
swirl-stabilized non-premixed flames with local extinction. Proc Combust Inst
2015;35(2):116774.
[17] Krieger G et al. Numerical simulation of oxy-fuel combustion for gas turbine
applications. Appl Therm Eng 2015;78:47181.
[18] Zhu X et al. Experimental study and RANS calculation on velocity and
temperature of a kerosene-fueled swirl laboratory combustor with and
without centerbody air injection. Int J Heat Mass Transf 2015;89:96476.
[19] Santhosh R, Basu S. Transitions and blowoff of unconfined non-premixed
swirling flame. Combust Flame 2015.
[20] Liu C et al. Investigations of the effects of spray characteristics on the flame
pattern and combustion stability of a swirl-cup combustor. Fuel
2015;139:52936.
[21] Mahesh S, Mishra D. Characterization of swirling CNG inverse jet flame in
recessed coaxial burner. Fuel 2015;161:18292.
[22] Bulat G, Jones W, Marquis A. NO and CO formation in an industrial gas-turbine
combustion chamber using LES with the Eulerian sub-grid PDF method.
Combust Flame 2014;161(7):180425.
[23] Khalil AE, Gupta AK. Towards distributed combustion for ultra low emission
using swirling and non-swirling flowfields. Appl Energy 2014;121:1329.
[24] Ghose P et al. Effect of air flow distribution on soot formation and radiative
heat transfer in a model liquid fuel spray combustor firing kerosene. Int J Heat
Mass Transf 2014;74:14355.
[25] Zheng Y et al. Large-eddy simulation of mixing and combustion in a premixed
swirling combustor with synthesis gases. Comput Fluids 2013;88:70214.
[26] Hermeth S et al. Bistable swirled flames and influence on flame transfer
functions. Combust Flame 2014;161(1):18496.
[27] Kim M-K et al. Effects of unstable flame structure and recirculation zones in a
swirl-stabilized dump combustor. Appl Therm Eng 2013;58(1):12535.
[28] Koyama M, Tachibana S. Technical applicability of low-swirl fuel nozzle for
liquid-fueled industrial gas turbine combustor. Fuel 2013;107:76676.
[29] Orbay R et al. Swirling turbulent flows in a combustion chamber with and
without heat release. Fuel 2013;104:13346.
[30] Steinberg A, Arndt C, Meier W. Parametric study of vortex structures and their
dynamics in swirl-stabilized combustion. Proc Combust Inst 2013;34
(2):311725.
[31] Dinesh KR et al. Swirl effects on external intermittency in turbulent jets. Int J
Heat Fluid Flow 2012;33(1):193206.
[32] De A, Acharya S. Parametric study of upstream flame propagation in hydrogenenriched premixed combustion: effects of swirl, geometry and premixedness.
Int J Hydrogen Energy 2012;37(19):1464968.
[33] Franzelli B et al. Large eddy simulation of combustion instabilities in a lean
partially premixed swirled flame. Combust Flame 2012;159(2):62137.
[34] Wolf P et al. Acoustic and large eddy simulation studies of azimuthal modes in
annular combustion chambers. Combust Flame 2012;159(11):3398413.
[35] Chmielewski M, Gieras M. Impact of variable geometry combustor on
performance and emissions from miniature gas turbine engine. J Energy Inst
2016.
[36] Saddawi SD, Kipouros T, Savill M. Computational engineering design for microscale combustion devices: a thermally improved configuration. In: ASME
Turbo Expo 2013: Turbine Technical Conference and Exposition. American
Society of Mechanical Engineers; 2013.
[37] Reddy VM, Sawant D, Kumar S. Studies on optimization of a liquid fuel based
low emission combustor. In: ASME 2012 Gas Turbine India
Conference. American Society of Mechanical Engineers; 2012.
[38] Hajitaheri S. Design optimization and combustion simulation of two gaseous
and liquid-fired combustors; 2012.
[39] Saddawi SD, Kipouros T, Savill M. Computational engineering design for microscale combustors. In: ASME Turbo Expo 2012: Turbine Technical Conference
and Exposition. American Society of Mechanical Engineers; 2012. p. 85969.

M.M. Torkzadeh et al. / Fuel 186 (2016) 734749


[40] Wankhede MJ, Bressloff NW, Keane AJ. Combustor design optimization using
co-kriging of steady and unsteady turbulent combustion. J Eng Gas Turbines
Power 2011;133(12):121504.
[41] Burmberger S, Sattelmayer T. Optimization of the aerodynamic flame
stabilization for fuel flexible gas turbine premix burners. J Eng Gas Turbines
Power 2011;133(10):101501.
[42] Motsamai OS, Snyman JA, Meyer JP. Optimization of gas turbine combustor
mixing for improved exit temperature profile. Heat Trans Eng 2010;31
(5):40218.
[43] Kulshreshtha DB, Channiwala S, Dikshit SB. Numerical simulation as design
optimization tool for gas turbine combustion chambers. In: ASME Turbo Expo
2010: Power for Land, Sea, and Air. American Society of Mechanical Engineers;
2010.
[44] Laraia M et al. A multi-objective design optimization strategy as applied to
pre-mixed pre-vaporized injection systems for low emission combustors.
Combust Theor Model 2010;14(2):20333.
[45] Fuligno L, Micheli D, Poloni C. An integrated approach for optimal design of
micro gas turbine combustors. J Therm Sci 2009;18(2):17384.
[46] Duchaine F, Morel T, Gicquel LM. Computational-fluid-dynamics-based kriging
optimization tool for aeronautical combustion chambers. AIAA J 2009;47
(3):63145.
[47] Nirmolo A, Woche H, Specht E. Temperature homogenization of reactive and
non-reactive flows after radial jet injections in confined cross-flow. Eng Appl
Comput Fluid Mech 2008;2(1):8594.
[48] Bhoi P, Channiwala S. Optimization of producer gas fired premixed burner.
Renew Energy 2008;33(6):120919.
[49] Iki N, Gruber A, Yoshida H. Anumerical and an experimental study for
optimization of a small annular combustor. In: Challenges of Power
Engineering and Environment. Springer; 2007. p. 142935.
[50] Janiga G, Thvenin D. Reducing the CO emissions in a laminar burner using
different numerical optimization methods. Proc Inst Mech Eng, Part A: J Power
Energy 2007;221(5):64755.
[51] Xiong Y et al. Multidisciplinary design optimization of aircraft combustor
structure: an industry application. AIAA J 2005;43(9):200814.
[52] Gordon R, Levy Y. Optimization of wall cooling in gas turbine combustor
through three-dimensional numerical simulation. J Eng Gas Turbines Power
2005;127(4):70423.
[53] Ganesan V, Kumar VJ. Gas turbine combustor: modelling and optimization. In:
ASME 2005 International Mechanical Engineering Congress and
Exposition. American Society of Mechanical Engineers; 2005.
[54] Menter FR et al. A correlation-based transition model using local variables
Part I: model formulation. J Turbomach 2006;128(3):41322.
[55] Dinesh KR et al. Influence of bluff-body and swirl on mixing and intermittency
of jets. Eng Appl Comput Fluid Mech 2010;4(3):37486.

749

[56] Pitsch H, Peters N. A consistent flamelet formulation for non-premixed


combustion considering differential diffusion effects. Combust Flame
1998;114(1):2640.
[57] Coelho P, Peters N. Unsteady modelling of a piloted methane/air jet flame
based on the Eulerian particle flamelet model. Combust Flame 2001;124
(3):44465.
[58] Poinsot T, Veynante D. Theoretical and numerical combustion. 3rd ed. RT
Edwards Inc.; 2012.
[59] 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):76174.
[60] Patel A, Kong S-C, Reitz RD. Development and validation of a reduced reaction
mechanism for HCCI engine simulations. SAE Technical Paper; 2004.
[61] Modest MF. Radiative heat transfer. Academic press; 2013.
[62] Burcat, garfield.chem.elte.hu/Burcat/THERM.DAT; 2014.
[63] Lin Y, Zhang HH. Component selection and smoothing in multivariate
nonparametric regression. Annals Stat 2006;34(5):227297.
[64] Deb K et al. A fast and elitist multiobjective genetic algorithm: NSGA-II. IEEE
Trans Evol Comput 2002;6(2):18297.
[65] Mattingly J, Heiser W, Pratt D. Aircraft engine design. 2nd ed. Reston, VA: AIAA
Education Series; 2002.
[66] Barlow R et al. Piloted methane/air jet flames: transport effects and aspects of
scalar structure. Combust Flame 2005;143(4):43349.
[67] Lefebvre AH. Gas turbine combustion. CRC Press; 1998.
[68] Kaddah K. Discharge coefficient and jet deflection studies for combustor liner
air-entry holes; 1964.
[69] Evans D, Noble M. Gas turbine combustor cooling by augmented backside
convection. J Eng Power 1979;101(1):10915.
[70] Masri A. Sydney swirl flow and dlame database Available from: <http://
sydney.edu.au/engineering/aeromech/thermofluids/swirl.htm>.
[71] Stein O, Kempf A. LES of the Sydney swirl flame series: a study of vortex
breakdown in isothermal and reacting flows. Proc Combust Inst 2007;31
(2):175563.
[72] Stein O, KEMPF AM, Janicka J. LES of the Sydney swirl flame series: an initial
investigation of the fluid dynamics. Combust Sci Technol 2007;179(1
2):17389.
[73] Hu L, Zhou L, Luo Y. Large-eddy simulation of the Sydney swirling
nonpremixed flame and validation of several subgrid-scale models. Num
Heat Trans, Part B: Funda 2008;53(1):3958.
[74] Yang Y, Kr SK. Large-eddy simulations of the non-reactive flow in the Sydney
swirl burner. Int J Heat Fluid Flow 2012;36:4757.
[75] Safavi M et al. A comparative study of turbulence models for swirling flow and
flame in gas turbine combustion chambers [in preparation].

Potrebbero piacerti anche