12
Modelling of non-premixed swirl burner flows using a Reynolds-stress turbulence closure A.E. German, T. Mahmud * Department of Chemical Engineering, School of Process, Environmental and Materials Engineering, University of Leeds, Leeds LS2 9JT, UK Received 30 March 2004; received in revised form 27 October 2004; accepted 27 October 2004 Available online 18 November 2004 Abstract In this computational study, the performance of a differential Reynolds-stress turbulence model has been assessed in predicting a turbulent, non-premixed combusting swirling flow of the type frequently found in practical combustion systems. Calculations are also performed using the widely employed eddy-viscosity based k3 turbulence model in order to examine the relative performances of these two closure models. The predictions are compared against the experimental data of mean axial and tangential velocities, turbulence quantities, gas temperatures and oxygen concentration collected in a 400 kW semi-industrial scale combustor fired with coke-oven gas using an industry-type swirl burner at the International Flame Research Foundation [17]. Computations of a corresponding non-combusting flow are also carried out and the predictions are compared with limited data available. The overall agreement between the measurements and the predictions obtained with both the k3 and Reynolds-stress turbulence models are reasonably good, in particular, the flame properties. However, some features of the isothermal and combusting flow fields, and the flame are better predicted by the Reynolds-stress model. The subcritical nature of the isothermal flow and the effects of combustion on the size and shape of the swirl-induced internal recirculation zone in the corresponding combusting flow are well simulated by this model. The k3 model fails to reproduce the subcritical nature of the isothermal flow. The predictions of this model erroneously show a general trend of the mean tangential velocity distribution to assume a forced-vortex profile. The levels of gas temperature and oxygen concentration in the internal recirculation zone and the enveloping shear region are on the whole better predicted by the Reynolds-stress model. q 2004 Elsevier Ltd. All rights reserved. Keywords: Combusting swirling flow modelling; Combustion modelling; Second-moment closure modelling 1. Introduction Swirling flows are widely used in industrial burners employed, for example, in power-station furnaces and gas- turbine combustors to provide stable and high-intensity flames. Flame structure and stability, and pollutant emis- sions strongly depend on the aerodynamic and mixing characteristics of the fuel and swirling combustion air jets in the near burner region. Over the last two decades, significant progress has been made in the development of compu- tational fluid dynamics (CFD) based models to simulate the performance of practical combustion systems. These models are now increasingly being used for the evaluation of the performance, and to assist in the design and development of such combustors. Reliable predictions of the combustion and pollutant formation processes occurring in the near burner region critically depend on the accuracy of the turbulent flow field calculation. The eddy-viscosity based k3 turbulence model is generally employed for flow calculations. The limitations of this model for predictions of both non-combusting (isothermal) and combusting strongly swirling flows, in particular, the size and strength of the swirl-induced internal recirculation zone (IRZ), are well known [1–4]. For improved predictions of the properties of swirling flows, a more detailed representation of the turbulent transport of momentum and scalar quantities is required. This can be achieved through the use of second- moment closures based on the solutions of modelled differential transport equations. Although more advanced methods, such as large-eddy simulation, are available, they 0016-2361/$ - see front matter q 2004 Elsevier Ltd. All rights reserved. doi:10.1016/j.fuel.2004.10.015 Fuel 84 (2005) 583–594 www.fuelfirst.com * Corresponding author. Tel.: C44 113 343 2431; fax: C44 113 343 2405. E-mail address: [email protected] (T. Mahmud).

Modelling of Non-premixed Swirl Burner Flows Using a Reynolds-stress Turbulence Closure

Embed Size (px)

DESCRIPTION

Free

Citation preview

  • ix

    ess

    n, T

    ntal a

    form

    e 18 N

    models are now increasingly being used for the evaluation turbulent transport of momentum and scalar quantities is

    Fuel 84 (2005)0016-2361/$ - see front matter q 2004 Elsevier Ltd. All rights reserved.In this computational study, the performance of a differential Reynolds-stress turbulence model has been assessed in predicting a turbulent,

    non-premixed combusting swirling flow of the type frequently found in practical combustion systems. Calculations are also performed using

    the widely employed eddy-viscosity based k3 turbulence model in order to examine the relative performances of these two closure models.

    The predictions are compared against the experimental data of mean axial and tangential velocities, turbulence quantities, gas temperatures

    and oxygen concentration collected in a 400 kW semi-industrial scale combustor fired with coke-oven gas using an industry-type swirl burner

    at the International Flame Research Foundation [17]. Computations of a corresponding non-combusting flow are also carried out and the

    predictions are compared with limited data available. The overall agreement between the measurements and the predictions obtained with

    both the k3 and Reynolds-stress turbulence models are reasonably good, in particular, the flame properties. However, some features of the

    isothermal and combusting flow fields, and the flame are better predicted by the Reynolds-stress model. The subcritical nature of the

    isothermal flow and the effects of combustion on the size and shape of the swirl-induced internal recirculation zone in the corresponding

    combusting flow are well simulated by this model. The k3 model fails to reproduce the subcritical nature of the isothermal flow. The

    predictions of this model erroneously show a general trend of the mean tangential velocity distribution to assume a forced-vortex profile. The

    levels of gas temperature and oxygen concentration in the internal recirculation zone and the enveloping shear region are on the whole better

    predicted by the Reynolds-stress model.

    q 2004 Elsevier Ltd. All rights reserved.

    Keywords: Combusting swirling flow modelling; Combustion modelling; Second-moment closure modelling

    1. Introduction

    Swirling flows are widely used in industrial burners

    employed, for example, in power-station furnaces and gas-

    turbine combustors to provide stable and high-intensity

    flames. Flame structure and stability, and pollutant emis-

    sions strongly depend on the aerodynamic and mixing

    characteristics of the fuel and swirling combustion air jets in

    the near burner region. Over the last two decades, significant

    progress has been made in the development of compu-

    tational fluid dynamics (CFD) based models to simulate the

    performance of practical combustion systems. These

    of the performance, and to assist in the design and

    development of such combustors. Reliable predictions of

    the combustion and pollutant formation processes occurring

    in the near burner region critically depend on the accuracy

    of the turbulent flow field calculation. The eddy-viscosity

    based k3 turbulence model is generally employed for flow

    calculations. The limitations of this model for predictions of

    both non-combusting (isothermal) and combusting strongly

    swirling flows, in particular, the size and strength of the

    swirl-induced internal recirculation zone (IRZ), are well

    known [14]. For improved predictions of the properties of

    swirling flows, a more detailed representation of theModelling of non-prem

    using a Reynolds-str

    A.E. Germa

    Department of Chemical Engineering, School of Process, Environme

    Received 30 March 2004; received in revised

    Available onlin

    Abstracted swirl burner flows

    turbulence closure

    . Mahmud*

    nd Materials Engineering, University of Leeds, Leeds LS2 9JT, UK

    27 October 2004; accepted 27 October 2004

    ovember 2004

    583594

    www.fuelfirst.commoment closures based on the solutions of modelled

    differential transport equations. Although more advanced

    methods, such as large-eddy simulation, are available, they

    doi:10.1016/j.fuel.2004.10.015

    * Corresponding author. Tel.: C44 113 343 2431; fax: C44 113 3432405.

    E-mail address: [email protected] (T. Mahmud).required. This can be achieved through the use of second-

  • the various sub-models embodied in the prediction pro-

    cedure, such as the turbulent combustion and radiation heat

    v r ~u ~k Z v mt v~k

    CP K r~3 (5)

    d / Ftransfer models. It is therefore important to validate the

    turbulence models directly against flow measurements.

    In the present computational study, the performance of a

    Reynolds-stress turbulence (RST) model, based on the

    modelled differential transport equations for stresses, has

    been evaluated against that of the standard k3 turbulence

    model [15] by simulating isothermal and combusting

    turbulent flows produced by an industry-type, non-premixed

    swirl burner with a divergent quarl. This study focuses on

    the issue of turbulence modelling. The widely used eddy-

    dissipation combustion model [16] has been adopted for the

    modelling of non-premixed combustion. The experiments

    were carried out in a semi-industrial scale combustor fired

    with coke-oven gas at the International Flame Research

    Foundation (IFRF), Holland [17]. The predictions obtained

    with the two turbulence models are compared against the

    experimental data of mean axial and tangential velocities,

    turbulence quantities, gas temperatures and oxygen

    concentration.

    2. Modelling of combusting flow

    Mathematical modelling procedure for turbulent com-

    busting flow involves numerical solutions of the time-mean

    conservation equations for mass, momentum, chemical

    species and thermal energy. Supplementary equations are

    solved to determine the turbulent momentum (Reynolds

    stresses) and scalar fluxes, the rate of combustion reaction,

    and the radiation heat transfer in the combustor. An

    overview of the aerodynamic and combustion calculation

    procedure employed in the present study is provided here.

    2.1. Conservation equations for fluid flow

    The Favre-averaged equations for conservation of mass

    and momentum for a steady, variable-density turbulent flowrequire prohibitively large computing resources even for

    relatively simple flows.

    In the past, substantial efforts were made on the

    development and validation of second-moment differential

    Reynolds-stress turbulence models for a range of flows (see

    reviews in Refs. [5,6]) including confined, isothermal

    swirling flows [3,711]. However, their evaluation for the

    calculation of combusting swirling flows, particularly in

    geometries relevant to industrial burner configurations, has

    been remarkably limited [1214]. Because of the scarcity of

    reliable mean velocity and turbulence data for combusting

    swirl burner flows, the performance of the turbulence

    models has generally been assessed indirectly against the

    gas temperature and species concentration measurements.

    This is not an entirely satisfactory approach as the

    predictions of such quantities are influenced strongly by

    A.E. German, T. Mahmu584can be expressed in concise forms in terms of Cartesianvxjj

    vxj sk vxj

    ~3-transport equation

    v

    vxj r ~uj ~3 Z v

    vxj

    mt

    s3

    v~3

    vxj

    CC31

    ~3~k

    P KC32r~32

    ~k(6)

    where P is the rate of production of turbulence energy and is

    given by:

    P ZKru00i u00jv ~uivxj

    (7)

    The model constants are assigned the following standard

    values [15]: CmZ0.09, C31Z1.44, C32Z1.92, skZ1.0 ands3Z1.3.

    RST model. The Reynolds-stress turbulence model

    employed in this study is based on the modelled partial

    differential equations for transport of the individual stressestensor notation as:

    v

    vxj r ~uj Z 0 (1)

    v

    vxjr ~ui ~uj

    Zv

    vxjm

    v ~uivxj

    Cv ~ujvxi

    K

    v P

    vxiK

    v

    vxj ru00i u00j

    (2)

    where ~ui and u00i are the Favre-averaged (density-weighted

    mean) and fluctuating velocity components respectively in

    the xi direction, P and r are the unweighted mean(conventional time-averaged) pressure and density of the

    mixture, and m is the laminar viscosity. The Reynolds

    stresses, ru00i u00j , are obtained using two different closuremodels: the eddy-viscosity based k3 model [15] and the

    RST model [5,18].

    2.2. Turbulence models

    k3 model. The Reynolds stresses are related to the rate

    of strain based on the Boussinesq hypothesis as:

    ru00i u00j ZKmtv ~uivxj

    Cv ~ujvxi

    C

    2

    3dij r ~k (3)

    The isotropic turbulent viscosity, mt, is given by the Prandtl

    Kolmogorov relation

    mt Z Cm r ~k2=~3 (4)

    where ~kZ 1=2u00i u00i is the turbulent kinetic energy, ~3 is therate of dissipation of ~k, and Cm is an empirical constant. Theturbulent kinetic energy and its dissipation rate are

    determined by solving their modelled transport equations.~k-transport equation:

    uel 84 (2005) 583594[5,18]. The transport equations for the Reynolds stresses,

  • strain redistribution, viscous dissipation and additional

    ud / Fproduction of the stresses.

    The convection and production terms are exact, while the

    remaining terms are modelled. The production term, Pij, can

    be expressed as:

    Pij ZKru00i u

    00k

    v ~ujvxk

    K ru00j u00kv ~uivxk

    (9)

    The diffusion term is modelled by a simple gradient-

    diffusion approximation [19] using the isotropic turbulent

    viscosity. The stress dissipation process is assumed to be

    isotropic and is modelled in terms of the rate of dissipation

    of turbulent kinetic energy as:

    ~3ij Z2

    3dij ~3 (10)

    The pressurestrain redistribution term, Pij, consists of

    two components: the return-to-isotropy term [20] and the

    Rapid or isotropization-of-production term [21]. It should

    be noted that the contribution of the wall reflection term,

    effect of which is confined in the proximity of the walls, is

    not included. Thus, the Pij term can be expressed as:

    Pij ZKC1 r~3~k

    u00i u00j K2

    3dij ~k

    KC2 r Pij K

    1

    3dijPkk

    (11)

    where C1 and C2 are model constants, and their values are

    taken [22] as 3.0 and 0.3, respectively.

    The Gij term represents the interaction between the mean

    pressure gradients and the density-weighted fluctuating

    velocities, and appears only in the density-weighted stress

    transport equations. The contribution of this term is small

    compared to other terms of Eq. (8) [23], and is neglected in

    the present calculation. The dissipation rate, ~3, is determinedby solving a transport equation compatible with the RST

    model and the turbulent kinetic energy, ~k, is obtaineddirectly from the normal stresses.

    2.3. Scalar transport equations

    The density-weighted transport equations for scalar

    quantities, such as chemical species and gas enthalpy, for

    a steady, turbulent flow can be written in general form using

    Cartesian tensor notation as:

    v r ~ui ~f ZK v ru00i f00 C Sf (12)ru00i u00j , are expressed in general form as:

    v

    vxk r ~uku00i u00j K

    v

    vxkmt

    vu00i u00jvxi

    !Z Pij CPij K r~3ij CGij

    (8)

    where the various terms from left to right represent,

    respectively, convection, diffusion, production, pressure

    A.E. German, T. Mahmvxi vxi2.4. Combustion model

    A suitable turbulent combustion model is required in

    order to determine the mean reaction rate of fuel, which

    allows the calculations of the source terms in the species and

    enthalpy transport equations. In the present calculation, the

    coke-oven gas is represented by a single chemical species in

    order to reduce the number of species transport equations.

    The measurements [17] show that the concentrations of

    carbon monoxide are rather small, in the range of 104

    105 ppm. Hence, the fuel is assumed to burn by a single-step

    chemical reaction to produce the final combustion products

    (carbon dioxide and water vapour)

    Fuel1 kg

    COxidants kg

    0Products1Cs kg

    (14)

    where s is the stoichiometric oxygen requirement, and is

    determined from the fuel composition given in Table 1.

    The turbulent non-premixed combustion process is

    simulated using the widely employed eddy-dissipation

    combustion model [16]. According to this model, the

    mean reaction rate of fuel is proportional to the inverse of

    the time-scale of the large-scale eddies characterised by the

    ratio ~k=~3, and to the smallest of the fuel, oxygen or productsconcentrations. The mean reaction rate is given by:

    Rfu Z A r~3

    min ~mfu;~mox

    ;B~mpr

    (15)where ~f and f 00 represent the Favre-averaged andfluctuating components of an instantaneous scalar quantity.

    In the present calculation, transport equations for the mass

    fraction of fuel ~mfu, oxidant ~mox and products ~mpr, andgas enthalpy ~h are solved. Sf stands for the time-averagedsource term representing the mean rate of formation or

    destruction Ri of a chemical species in the speciestransport equations, and the rate of heat generation by

    combustion (DH Rfu, where DH is the calorific value of thefuel) and net heat gained due to thermal radiation QR in theenthalpy transport equation.

    The turbulent scalar fluxes, ru00i f00 , can be determined bytwo different types of closure models: a gradient-diffusion

    model based on the eddy-viscosity approach and a second-

    moment scalar flux model based on the transport equations

    for scalar flux components analogous to Eq. (8). The latter

    model leads to a significant increase in computing time and

    storage requirement. Hence in the present calculation, in

    common with previous studies [13,14], the turbulent scalar

    fluxes are modelled using the gradient-diffusion approach

    [15] as

    ru00i f00 ZKmt

    st

    v ~f

    vxi(13)

    where st(Z0.7) stands for the turbulent Prandtl or Schmidtnumber.

    uel 84 (2005) 583594 585~k s 1 Cs

  • 2.5. Thermal radiation model

    2.6. Numerical solution procedure

    velocities is used. In order to improve numerical stability

    when the RST model is employed, a staggered arrangement

    d / FThe computational procedure for combusting swirling

    flow is based on finite-volume numerical solutions of theThe radiation heat transfer is simulated by means of the

    non-equilibrium diffusion radiation model [24]. This model,

    which has been used before for the simulation of pulverised-

    coal [25,26] and natural gas [27] flames, is easy to apply,

    computationally efficient and compatible with the finite-

    volume solution method of the governing transport

    equations. The radiative transfer equation for an absorbing

    and emitting medium is expressed as:

    v

    vxi

    1

    3k

    vT4Rvxi

    Z kT4R K ~T4 (16)

    where T4R and ~Tare the radiation and mean gas temperatures,respectively, and k is the gas absorption coefficient. The

    above transfer equation is solved for T4R which allows the

    calculation of the radiation source term, QR, in the enthalpytransport equation:

    QR Z ksT4R K ~T4 (17)where s is the StefenBoltzman constant.where A and B are constants which take the values of 4 and

    0.5, respectively.

    Table 1

    Fuel composition and the burner operating conditions

    Coke-oven gas analysis (vol.%) Burner operating conditions

    CH4 22.4 Fuel flow rate (kg/h) 38.4

    Higher hydrocarbons 3.6 Air flow rate (kg/h) 1200

    H2 62.6 Air temperature (K) 300

    CO 5.5 Thermal input (kW) 400

    CO2 1.2 Air swirl number 1.4

    O2 0.2 Air velocity (m/s) 19.2

    N2 4.5 Fuel velocity (m/s) 33.6

    DensityZ0.4142 kg/m3; calorific valueZ40 MJ/kg; stoichiometric airZ13.4 kg/kg of fuel.

    A.E. German, T. Mahmu586governing transport equations for the mean axial ~U, radial ~V and tangential ~W velocities; ~k, ~3 and six components ofthe Reynolds stress (u002 , v002 , w002 , u00v00 , u00w00 , v00w00); meanscalar quantities ( ~mfu, ~mox, ~mpr, ~h); and the radiation heattransfer. The combustor simulated in this study is

    cylindrical and is fired along its axis, which allows the

    flow to be modelled as two-dimensional and axisymmetric.

    The basic CFD code adapted for the calculations has

    been described in a number of references, see, for example,

    Refs. [25,26]. This code is based on the k3 turbulence

    model and the hybrid (central/upwind) and QUICK [28]

    differencing schemes. In the present study, the RST model

    and a higher-order boundedness preserving differencingfor the shear stresses is adopted as suggested in Ref. [30].

    The rest of the variables are stored at the scalar grid nodes. In

    recirculating flow calculation, numerical diffusion errors

    resulting from the use of a first-order convective differencing

    scheme may be significant when the flow-to-grid skewness is

    large. The use of a higher-order scheme such as the QUICK

    scheme [28] can minimise numerical diffusion. However,

    this scheme does not possess the boundedness property and

    may generate physically unrealistic solutions. Consequently,

    the convective terms of the transport equations have been

    approximated by a third-order boundedness preserving

    scheme, known as the curvature-compensated convective

    transport (CCCT) algorithm [29]. The diffusion terms are

    discretised by the central-difference scheme.

    The solutions of the elliptic transport equations require

    specification of boundary conditions on the four sides of the

    computational domain. The inlet boundary conditions

    employed in the computations are described below. At the

    symmetry axis, the values of the radial and tangential

    velocities, shear stresses, and the cross-stream gradients of

    all other variables are set to zero. No-slip boundary

    conditions and log-law based wall functions are applied

    along the solid walls. Diffusion of u00i u00j normal to the wall isset to zero. At the outlet boundary, the zero streamwise

    gradient conditions are imposed for all variables except for

    the axial velocity, which is adjusted to satisfy the overall

    mass continuity.

    Finally, the discretised momentum equations coupled

    with a Poisson-type pressure correction equation are solved

    iteratively for the velocity components and pressure using

    the PISO algorithm [31]. Solution of the discretised

    equations for all other variables and updating of the

    physical properties are also performed in each iteration.

    The discretised equations are solved using the tridiagonal

    matrix algorithm (TDMA).

    3. Application of the model

    3.1. The experimental case

    The combustor used at IFRF [17] was cylindrical with an

    internal diameter of 0.44 m and a length of 2.0 m, and was

    fired by an industry-type swirl burner. The burner consisted

    of two concentric nozzles with a cylindrical bluff bodyscheme [29] have been incorporated into the original code,

    and the validation results are presented in this paper.

    The transport equations (presented in Cartesian tensor

    notation in the preceding sections) in their two-dimensional,

    axisymmetric form in cylindrical coordinates are discretised

    by integrating over control volumes covering one half-plane

    of the flow domain. A conventional staggered arrangement of

    the control volumes associated with the axial and radial

    uel 84 (2005) 583594inserted in the primary nozzle and fitted with a divergent

  • of the

    A.E. German, T. Mahmud / Fuel 84 (2005) 583594 587quarl of 208 half-angle. This combustor is illustratedschematically in Fig. 1. The fuel (coke-oven gas) was

    introduced through the primary nozzle and the combustion

    air through the secondary nozzle. Swirl in the combustion

    air was imparted by a tangential-entry swirl generator

    located upstream of the burner throat which produced solid-

    body rotation flows in the secondary nozzle. The burner

    quarl and the cylindrical walls were refractory lined. The

    confinement ratio, defined as the ratio of the cylindrical

    chamber diameter to the quarl exit diameter, was 1.15.

    The semi-industrial swirling flame simulated in this

    study is of 400 kW thermal input. The coke-oven gas

    composition and the operating conditions of the burner are

    given in Table 1. In the experiment, a short and intense non-

    premixed swirling flame in the vicinity of the burner quarl

    was produced. The fuel jet did not penetrate into the IRZ as

    it was rapidly entrained by the swirling air stream and

    combustion occurred on the boundary of the IRZ [17].

    According to the IFRF classification, this flame is referred to

    as type-2 non-premixed flame. Measurements of the mean

    axial and tangential velocities, and the components of

    normal stresses (u002 and w002) were obtained using aLDA system. Gas temperatures were measured using

    Fig. 1. Geometrya conventional intrusive suction pyrometer and species

    concentrations by analysing gas samples obtained using a

    water quenched probe. The measurement stations were

    Fig. 2. Computational grlocated at 0.19, 0.373, 0.543 and 1.623 m from the burner

    inlet.

    3.2. Computational details

    The calculations were carried out using the measured

    inlet flow conditions where available. The measured profiles

    of mean axial and tangential velocity, and the normal

    stresses (u002 and w002) at the quarl inlet were employed asthe inlet boundary conditions. In the absence of experimen-

    tal data, the mean radial velocity was set to zero and the

    normal stress v002 was assumed [32] to be 0:5u002 . The threecomponents of shear stresses (u00v00 , u00w00 , v00w00) were takento be zero. The measured profiles of normal stresses at the

    quarl inlet were used to calculate the inlet turbulent kinetic

    energy ~kin. The inlet energy dissipation rate ~3 wasestimated from ~k

    1:5in =[, with a constant mixing-length, [,

    taken as 0.33 times the nozzle dimension (annular widths of

    the primary and secondary nozzles). As mentioned above,

    the IFRF combustor walls were refractory lined. The heat

    losses through the walls were less than 5% of the thermal

    input [17]. Consequently, the combustor was treated as

    adiabatic.

    IFRF combustor.The computations were performed on a 54 (axial)!30(radial) grid. The computational grid in the near burner

    region is shown in Fig. 2. The grid lines were uniformly

    id near the burner.

  • in Figs. 3 and 4, respectively. For the isothermal flow, the

    k3 model predicts a swirl-induced closed IRZ anchored to

    rns us

    A.E. German, T. Mahmud / Fuel 84 (2005) 583594588the bluff body and a forward flow in the downstream region,

    which approaches the fully developed conditions near the

    exit of the combustor. In contrast, the RST model predicts a

    long IRZ, surrounded by an annular forward flow, extending

    to the exit of the combustor. The high confinement of thespaced in the radial direction and expanded in the axial

    direction, downstream of the burner quarl. The inclined wall

    of the quarl was represented by a series of steps.

    Calculations were also performed on a 32!24 grid withvirtually identical results.

    4. Results and discussion

    4.1. Effect of combustion on the flow pattern

    The predicted combusting and corresponding flow

    patterns in the form of non-combusting (isothermal)

    velocity vectors using the k3 and RST models are shown

    Fig. 3. Predicted combusting flow pattecylindrical chamber suppresses flow separation at the quart

    Fig. 4. Predicted isothermal flow patterns usexit, and as a consequence a very small reverse flow is

    predicted at the corner by both models of turbulence. The

    predicted and measured boundaries of the IRZ for

    isothermal and combusting flows are compared in Fig. 5.

    The isothermal flow measurements show a long IRZ

    extending up to the combustor exit without closing, which

    reveals the subcritical nature of this strongly swirling flow

    [17]. This is correctly captured by the RST model, whereas

    the k3 model fails to reproduce the basic feature of the

    flow. The failure of the k3 model in predicting subcritical

    features of cold swirling flows has also been observed in

    previous calculations [3,8,10].

    Comparison of the experimental data [17] presented in

    Fig. 5(a) and (b) reveals the effects of combustion on the flow

    pattern. The main effects are the reduction of both the size

    and the strength of the IRZ. The subcritical flow in the

    isothermal condition becomes supercritical when combus-

    tion takes place, resulting in the formation of a smaller and

    closed IRZ. Figs. 3 and 5(b) show that for the combusting

    flow both the k3 and RST models predict a closed IRZ in

    agreement with the experimental data. However, the RST

    ing (a) k3 model and (b) RST model.model predictions correctly show the forward flow near

    ing (a) k3 model and (b) RST model.

  • rmal and (b) combusting flow (C, experimental; - - -, k3 model; , RST model).

    ud / Fuel 84 (2005) 583594 589the axis of the combustor at xZ0.543 m (where x is the axialdistance measured from the burner throat). The reduction of

    the size and strength of the IRZ in the combusting flow is

    due to the decrease of the level of swirl in the combustor.

    The measurements show that the axial velocities in

    the forward flow region increase significantly due to

    combustion-induced flow acceleration while the tangential

    velocities are slightly altered. Consequently, the ratio of the

    tangential to axial momentum fluxes decreases substantially,

    resulting in a marked reduction of the inlet swirl number in

    the near burner region. The swirl number is defined as the

    ratio of the tangential momentum flux to axial momentum

    flux, as:

    Sn Z

    R2R1

    r ~U ~Wr2 dr=R2

    R2R1

    r ~U2r dr (18)

    where R1 and R2 are the inner and outer radii of the

    annular duct of the combustion air for the burner inlet

    swirl number, and for the swirl number at a downstream

    Fig. 5. Comparisons of measured and predicted IRZ boundaries for (a) isothe

    A.E. German, T. Mahmlocation, R1 is taken as zero and R2 is the quarl/combustor

    wall radius.

    The variations of the swirl numbers in the combusting

    flow, obtained using the RST model calculated and

    measured velocity distributions, along the length of the

    combustor are shown in Fig. 6. The values of the swirl

    number calculated using the k3 model predicted velocities

    are similar to those based on the RST model. As can be seen,

    the inlet swirl number of 1.4 is reduced to about 0.3 within

    the burner quarl and then it slowly increases in the

    downstream region where an experimental value of about

    0.5 and calculated 0.4 are reached. The values of the swirl

    number in the near burner region are very close to the

    minimum swirl level required in the experiment to establish

    a reverse flow [17]. It is important to note that measure-

    ments of a range of combusting swirling flows carried out in

    the IFRF combustor [17] reveals that the effects of

    combustion on the IRZ properties depend on the extent of

    reduction of the initial swirl level, which in turn depends onthe location of the flame front and the degree of flow

    acceleration.

    4.2. Comparison of predicted and measured flow fields

    The predicted and measured radial profiles of mean axial

    and tangential velocity for the combusting flow at four

    stations are shown in Figs. 7 and 8, respectively. The first

    station, xZ0.19 m, is within the burner quarl and the rest ofthe stations, xZ0.373, 0.543 and 1.623 m, are in thecylindrical chamber. Predictions are shown for the k3 and

    RST models. At the first station (xZ0.19 m), significantdiscrepancies exist between the predicted and measured

    axial and tangential velocities. The predicted widths of the

    IRZ are much too small compared to the measurement

    resulting in a wider shear layer enveloping this zone.

    Consequently, the steep gradients of the measured velocity

    profiles are not reproduced in the predictions. The predicted

    flow development immediately downstream of the burnerFig. 6. Comparisons of measured and predicted swirl number (6,

    experimental; , RST model).

  • d / FA.E. German, T. Mahmu590inlet is strongly influenced by the specified inlet values of

    the radial velocity and energy dissipation rate (3), as

    demonstrated in previous computational studies [33,34].

    The discrepancies between the predictions and measure-

    ments at this station are believed to be due to the use of zero

    radial velocity and 3 estimated from an assumed mixing-

    length as the inlet boundary conditions in the calculations.

    At the second station (xZ0.373 m), the axial velocitiespredicted by the k3 and RST models are in reasonably good

    agreement with the experimental data. However, differences

    between the predictions of two turbulence models are

    evident in the IRZ. It is difficult to draw a definite

    Fig. 7. Comparisons of measured and predicted mean axial velocity profiles

    (6, experimental; - - -, k3 model; , RST model).uel 84 (2005) 583594conclusion with regard to their performance in this region.

    As can be seen in Fig. 8, the tangential velocities are

    significantly underpredicted by both the turbulence models

    at this station. At the third station (xZ0.543 m), bothmodels of turbulence fail to reproduce the measured trend of

    the axial velocity profile in the forward flow region. It is

    interesting to note that at this station, the shape of the

    measured axial velocity profile resembles that of an

    unconfined flow and the measured tangential velocity

    profile shows an unexpected peak near the wall. However,

    the axial and tangential velocities predicted by the RST

    Fig. 8. Comparisons of measured and predicted mean tangential velocity

    profiles (6, experimental; - - -, k3 model; , RST model).

  • response to the swirl.

    the neglect in the calculation of the effect of soot on

    the radiation heat transfer from the flame. The production of

    soot in the near burner region influences the local gas

    temperature through its effect on the absorption coefficient

    of the medium. A previous modelling study [2] of swirling

    natural gas flame has demonstrated the smoothing effect of

    radiation from soot on the temperature distributions. The

    formation and oxidation of soot is a complex process and no

    attempt has been made in the present study to include this in

    Fig. 9. Comparisons of measured and predicted axial turbulence intensity

    profiles (6, experimental; - - -, k3 model; , RST model).

    ud / FIn the absence of detailed measurements of turbulence

    quantities, the predicted levels of axial,u002

    p,

    and tangential,w002

    p, turbulence intensities obtained

    using the k3 and RST models are compared against the

    data in Figs. 9 and 10, respectively. As can be seen, the RST

    model predicted turbulence intensities are lower and

    generally in better agreement with the measurements

    compared to those obtained from the k3 model. Previous

    studies [8,9] also reported that the k3 model overestimated

    the levels of stresses in confined, isothermal swirling flows,

    while the RST model correctly returned reduced levels of

    stresses in response to the swirl.

    4.3. Comparison of predicted and measured

    flame properties

    The predicted radial profiles of gas temperature based on

    the flow calculations using the k3 and RST models are

    compared with the measurements in Fig. 11. As can be seen,

    the general features of the measured temperature profiles are

    reasonably well predicted by both the turbulence models. At

    the first station (xZ0.19 m) within the burner quarl, thepredictions are in good agreement with the data in the

    central region inside the IRZ. It should be noted that at this

    station the width of the IRZ is significantly underpredicted

    by both the models of turbulence (see Fig. 7). The peaks in

    the predicted temperature profiles at a radial distance (r) of

    about 0.1 m, reveal that combustion occurs in the shear layer

    surrounding the IRZ, which is a typical feature of type-2

    non-premixed flames [17]. Although this is not evident from

    the measured temperature profile, the measured carbon

    dioxide concentration distribution (not presented here)

    shows a peak near the quarl wall. The predicted highmodel near the axis of the combustor are in better agreement

    with the data compared to that of the k3 model predictions.

    At the last station (xZ1.623 m), both the k3 and RSTmodels, in agreement with the data, predict flat axial

    velocity profiles. The predicted tangential velocity distri-

    bution using the RST model at this station is in much better

    agreement with the data compared to that predicted by the

    k3 model. The latter model erroneously predicts a forced-

    vortex profile. The k3 model predictions in general show a

    tendency of the tangential velocity distribution to approach

    the forced-vortex profile, which has also been reported by

    previous investigators for computations of isothermal [1,8]

    and combusting [2] flows. This has been attributed to

    deficiencies in the 3-transport equation in Ref. [1] and to the

    overestimation of the diffusive transport of momentum in

    the radial direction by the k3 model in Ref. [8]. The latter

    has also been revealed by the predicted levels of shear

    stresses (not shown here) by the two turbulence models in

    our calculations. The comparison clearly shows the

    reduction in stress levels produced by the RST model in

    A.E. German, T. Mahmtemperatures in the shear layer are presumably due touel 84 (2005) 583594 591the radiation model. Downstream of the quarl, at xZ0.373

  • A.E. German, T. Mahmud / F592and 0.543 m, the predictions are again in good agreement

    with the data. At these stations, the peaks in the measured

    temperature profiles are evident near the combustor wall. At

    xZ0.375 m, although the predicted value of the temperaturepeak is in good agreement with the data, its radial location is

    further away from the wall. At the last station (xZ1.623 m),which is located in the post-flame region far downstream of

    the IRZ, both models underpredict temperatures by about

    250 8C, probably due to uncertainty in the wall heat transfer

    boundary conditions employed in the calculation. In

    general, the predicted levels of temperature obtained with

    Fig. 10. Comparisons of measured and predicted tangential turbulence

    intensity profiles (6, experimental; - - -, k3 model; , RST model).uel 84 (2005) 583594the k3 and RST models are qualitatively similar (with

    a maximum difference of about 175 8C) although in the IRZand the enveloping shear region the latter model predictions

    are in better agreement with the measurements.

    Comparison between the predicted and measured radial

    profiles of oxygen concentration is shown in Fig. 12. At the

    first two stations (xZ0.19 and 0.373 m), both modelspredict, in agreement with the measurements, virtually zero

    oxygen concentration in the IRZ. In the region outside the

    IRZ dominated by the combustion air flow (see Fig. 3), the

    predicted and measured oxygen concentrations increase

    to a maximum value near the wall. As can be seen,

    Fig. 11. Comparisons of measured and predicted gas temperature profiles

    (6, experimental; - - -, k3 model; RST model).

  • A.E. German, T. Mahmud / Fthe predictions obtained from both the turbulence models

    are identical at the first station, and at xZ0.373, the RSTmodel returns slightly better predictions. At the third station

    (xZ0.543 m), the k3 model significantly overestimatesoxygen concentration in the IRZ whereas the RST model

    predictions are in better agreement with the data. At the last

    station (xZ1.623 m), both models overpredict oxygenconcentration near the axis of the combustor, but the

    predictions are in good agreement with the experimental

    Fig. 12. Comparisons of measured and predicted oxygen concentration

    profiles (6, experimental; - - -, k3 model; , RST model).5. Concluding remarks

    The performance of the standard eddy-viscosity based

    k3 and RST turbulence models in predicting combusting

    swirling flow has been assessed against the LDA flow and

    combustion data [17], collected in a gas fired semi-industrial

    combustor at IFRF. Calculations of a corresponding non-

    combusting (isothermal) flow are also performed in order to

    examine the effect of combustion on the flow field and the

    predictions are compared with a limited amount of data

    available.

    Although the k3 turbulence model reasonably well

    predicts the overall flow and combustion characteristics in

    the combustor, some features of both the isothermal and

    combusting flow fields, and the flame are better predicted by

    the RST model. The subcritical nature of the isothermal flow

    and the effects of combustion on the size and shape of the

    swirl-induced IRZ in the corresponding combusting flow are

    well simulated by the RST model. The k3 model fails to

    reproduce the subcritical nature of the isothermal flow. The

    predictions of this model erroneously show a general trend

    of the mean tangential velocity distribution to assume a

    forced-vortex profile. The predicted turbulence intensities

    using the RST model are generally in better agreement with

    the measurements compared to those obtained from the k3

    model. The levels of gas temperature and oxygen concen-

    tration in the IRZ and the enveloping shear region are on thedata further away from the axis (rO0.08 m). In general, thepredicted oxygen concentration profile at this station is

    relatively flat compared to the measured one presumably

    due to the overestimation of the diffusive transport of mass.

    The use of a second-moment closure for the calculation of

    the turbulent scalar fluxes in the transport equations

    (Eq. (12)) would probably enhance the quality of predic-

    tions. On the whole, the prediction of oxygen concentration

    distributions with the RST model is better than that given by

    the k3 model.

    In general, the predictions of the swirling flow field and

    flame properties in the present combustor obtained with the

    RST model are in better agreement with the data compared

    to that of the standard k3 model predictions. However, for

    the present flow with a high inlet swirl number of 1.4, it is

    expected that the differences between the predictions of the

    k3 and RST models should be greater than that displayed

    in previous figures. This, unexpected, performance of the

    turbulence models may be explained in terms of the effect

    of combustion on the swirl level. As revealed in Fig. 6,

    there is a drastic reduction of the initial level of swirl from

    1.4 to about 0.3 in the near burner region resulting from the

    increase of the axial momentum due to combustion-

    induced flow acceleration. Consequently, the effects of

    swirl on the mean flow and turbulence fields are

    significantly reduced.

    uel 84 (2005) 583594 593whole better predicted by the RST model.

  • For the type-2 non-premixed flame considered in this

    study, the combustion front is located in the vicinity of

    the burner as observed experimentally [17], which drasti-

    cally reduces the swirl number from 1.4 at the burner inlet to

    about 0.3 in this region due to the increase in axial

    momentum. As a consequence, the effect of swirl on the

    flow field is significantly reduced. The experiments carried

    out at IFRF demonstrated that the effect of combustion on

    the swirling flow field depends on the location of the flame

    front and the degree of flow acceleration. This suggests that

    the difference between the performances of these two

    turbulence models will depend on the flame types.

    Acknowledgements

    The financial support provided by Lagoven S. A. of

    Venezuela to A. German to undertake this research is

    gratefully acknowledged.

    References

    [5] Launder BE. Int J Heat Fluid Flow 1989;10:282.

    [6] Hanjalic K. Int J Heat Fluid Flow 1994;15:178.

    [7] Nikjooy M, So RMC. Int J Numer Meth Eng 1989;28:861.

    [8] Hogg S, Leschziner MA. J AIAA 1989;27:57.

    [9] Jones WP, Pascau A. J Fluids Eng, Trans ASME 1989;111:248.

    [10] Weber R, Visser BM, Boyan F. Int J Heat Fluid Flow 1990;10:225.

    [11] Nikjooy M, Mongia HC. Int J Heat Fluid Flow 1991;12:12.

    [12] Hogg S, Leschziner MA. In: Proceedings of the third international

    conference on computational combustion, Antibes, France 1989.

    [13] Lockwood FC, Shen B. Proc Combust Inst 1994;25:503.

    [14] Landenfeld T, Kremer A, Hassel EP, Janicka J. In: Proceedings of the

    eleventh symposium on turbulent shear flows, Grenoble 1997.

    [15] Launder BE, Spalding DB. Comput Meth Appl Mech Eng 1974;

    3:269.

    [16] Magnussen BF, Hjertager BH. Proc Combust Inst 1978;17:719.

    [17] Weber R, Dugue J. Prog Energy Combust Sci 1992;18:349.

    [18] Launder BE, Reece GJ, Rodi W. J Fluid Mech 1975;68:537.

    [19] Shir CC. J Atmos Sci 1973;30:1327.

    [20] Rotta JC. Z Phys 1951;29:547.

    [21] Naot D, Shavit A, Woifshtein M. Israel J Technol 1970;8:259.

    [22] Gibson MM, Younis BA. Phys Fluids 1986;29:38.

    [23] Jones WP. In: Kolimann W, editor. Prediction methods for turbulent

    flows. London: Hemisphere; 1980. p. 379.

    [24] Arscott JA, Gibb J, Jenner R. In: Weinberg FJ, editor. Proceedings of

    the European symposium. Sheffield, UK: The Combustion Institute;

    1973. p. 675.

    [25] Lockwood FC, Mahmud T. Proc Combust Inst 1988;22:165.

    A.E. German, T. Mahmud / Fuel 84 (2005) 583594594[1] Sloan DG, Srnith PJ, Smoot LD. Prog Energy Combust Sci 1986;12:

    163.

    [2] Boardman RD, Eatough CN, Germane GJ, Smoot LD. Combust Sci

    Technol 1993;93:193.

    [3] Benim AC, Nahavandi A. In: Hanjalic K, Nagano Y, Tummers M,

    editors. Proceedings of turbulence, heat and mass transfer 4, 2003.

    p. 715.

    [4] Van Maele K, Merci B, Dick E. In: Hanjalic K, Nagano Y,

    Tummers M, editors. Proceedings of turbulence, heat and mass

    transfer 4, 2003. p. 931.[26] Lockwood FC, Mahmud T, Yehia MA. Fuel 1998;77(12):1329.

    [27] Costa M, Costen P, Lockwood FC, Mahmud T. Proc Combust Inst

    1990;23:973.

    [28] Leonard BP. Comput Meth Appl Mech Eng 1979;19:59.

    [29] Gaskell P, Lau A. Int J Numer Meth Fluids 1988;8:617.

    [30] Pope SB, Whitelaw JH. J Fluid Mech 1976;73:9.

    [31] Issa RI. J Comput Phys 1986;62:40.

    [32] Habib MA, Whitelaw JH. Numer Heat Transfer 1982;5:145.

    [33] Sturgess GJ, Syed SA, McManus KR. AIAA paper 83 1983 p. 1263.

    [34] Leschziner MA, Rodi W. J AIAA 1984;22:1742.

    Modelling of non-premixed swirl burner flows using a Reynolds-stress turbulence closureIntroductionModelling of combusting flowConservation equations for fluid flowTurbulence modelsScalar transport equationsCombustion modelThermal radiation modelNumerical solution procedure

    Application of the modelThe experimental caseComputational details

    Results and discussionEffect of combustion on the flow patternComparison of predicted and measured flow fieldsComparison of predicted and measured flame properties

    Concluding remarksAcknowledgementsReferences