• Aucun résultat trouvé

The importance of thermal radiation transfer in laminar diffusion flames at normal and microgravit

N/A
N/A
Protected

Academic year: 2021

Partager "The importance of thermal radiation transfer in laminar diffusion flames at normal and microgravit"

Copied!
10
0
0

Texte intégral

(1)

Publisher’s version / Version de l'éditeur:

Journal of Quantitative Spectroscopy and Radiative Transfer, 112, 7, pp.

1241-1249, 2011

READ THESE TERMS AND CONDITIONS CAREFULLY BEFORE USING THIS WEBSITE. https://nrc-publications.canada.ca/eng/copyright

Vous avez des questions? Nous pouvons vous aider. Pour communiquer directement avec un auteur, consultez la première page de la revue dans laquelle son article a été publié afin de trouver ses coordonnées. Si vous n’arrivez pas à les repérer, communiquez avec nous à PublicationsArchive-ArchivesPublications@nrc-cnrc.gc.ca.

Questions? Contact the NRC Publications Archive team at

PublicationsArchive-ArchivesPublications@nrc-cnrc.gc.ca. If you wish to email the authors directly, please see the first page of the publication for their contact information.

NRC Publications Archive

Archives des publications du CNRC

This publication could be one of several versions: author’s original, accepted manuscript or the publisher’s version. / La version de cette publication peut être l’une des suivantes : la version prépublication de l’auteur, la version acceptée du manuscrit ou la version de l’éditeur.

For the publisher’s version, please access the DOI link below./ Pour consulter la version de l’éditeur, utilisez le lien DOI ci-dessous.

https://doi.org/10.1016/j.jqsrt.2010.08.021

Access and use of this website and the material on it are subject to the Terms and Conditions set forth at

The importance of thermal radiation transfer in laminar diffusion flames

at normal and microgravit

Liu, Fengshan; Smallwood, Gregory J.; Kong, Wenjun

https://publications-cnrc.canada.ca/fra/droits

L’accès à ce site Web et l’utilisation de son contenu sont assujettis aux conditions présentées dans le site LISEZ CES CONDITIONS ATTENTIVEMENT AVANT D’UTILISER CE SITE WEB.

NRC Publications Record / Notice d'Archives des publications de CNRC:

https://nrc-publications.canada.ca/eng/view/object/?id=d7189bb1-e4e2-4a17-8097-5ae95fb362d0

https://publications-cnrc.canada.ca/fra/voir/objet/?id=d7189bb1-e4e2-4a17-8097-5ae95fb362d0

(2)

The importance of thermal radiation transfer in laminar diffusion

flames at normal and microgravity

Fengshan Liu

a,n

, Gregory J. Smallwood

a

, Wenjun Kong

b

a

National Research Council 1200 Montreal Road, Ottawa, Ontario, Canada K1A 0R6 b

Chinese Academy of Sciences, PO Box 2706, Beijing 100190, China

a r t i c l e

i n f o

Available online 8 September 2010

Keywords:

Laminar diffusion flame Soot formation Non-grey radiation Microgravity

a b s t r a c t

The importance of radiation heat loss in laminar and turbulent diffusion flames at normal gravity has been relatively well recognized in recent years. There is currently lack of quantitative understanding on the importance of radiation heat loss in relatively small scale laminar diffusion flames at microgravity. The effects of radiation heat transfer and radiation absorption on the structure and soot formation characteristics of a coflow laminar ethylene/air diffusion flame at normal- and microgravity were numerically investigated. Numerical calculations were conducted using GRI-Mech 3.0 combustion chemistry without the NOxmechanism and complex thermal and transport

properties, an acetylene based soot formation model, and a statistical narrow-band correlated-k non-grey gas radiation model. Radiation heat transfer and radiation absorption in the microgravity flame were found to be much more important than their counterparts at normal gravity. It is important to calculate thermal radiation transfer accurately in diffusion flame modelling under microgravity conditions.

Crown Copyright & 2010 Published by Elsevier Ltd. All rights reserved.

1. Introduction

Thermal radiation is often the dominant heat transfer mode in many systems involving large scale and high temperature, such as furnaces and combustors, and fire spread. The importance of radiation transfer in combus-tion and flame processes has drawn increased attencombus-tion over the last few decades. Even in relatively small sized laminar discussion flames radiation heat transfer can still be important to influence various flame processes, such as extinction at low stretch[1]and soot formation[2,3]. This is because radiation heat transfer affects the flame temperature, which in turn alters various kinetically controlled phenomena due to the strong temperature dependence of such processes. The strong coupling between radiation and soot kinetics in sooting flames at

earth gravity was first recognized by Kaplan et al.[4]and Sivathanu and Gore[5] in their numerical studies using crude combustion chemistry based on either the one-step reaction model or the state relationship. The importance of radiation heat transfer in laminar diffusion flames at earth gravity and atmospheric pressure was later numerically investigated in several studies by Liu et al.[2,3,6], using detailed combustion chemistry, a semi-empirical (acet-ylene based) soot model, and the statistical narrow-band correlated-k (SNBCK) non-grey gas radiation model. These studies revealed that even in such fairly small diffusion flames thermal radiation loss can still be important and should be accounted for in numerical modelling of laminar flames, especially for the 64 mm tall laminar coflow ethylene/air diffusion flame[2]and the oxygen enriched ethylene counterflow diffusion flame [3], where soot loading is relatively high. On the other hand, however, radiation absorption can be important or relatively unimportant, depending on flame configuration (counter-flow or co(counter-flow), flame size, and soot loading. The strong Contents lists available atScienceDirect

journal homepage:www.elsevier.com/locate/jqsrt

Journal of Quantitative Spectroscopy &

Radiative Transfer

0022-4073/$ - see front matter Crown Copyright & 2010 Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.jqsrt.2010.08.021

n

Corresponding author.

(3)

coupling between flame temperature, soot formation, and NOxemission and thermal radiation due to gas and soot in propane-fuelled oxygen-enriched turbulent jet diffusion flames was more recently investigated by Wang et al.[7]

using advanced soot formation and non-grey gas radiation models. Their study once again found that accurate modelling of thermal radiation due to both non-grey gas and soot is very important for accurate predictions of soot and NOx formation. It is worth pointing out that it is essential to employ an accurate non-grey gas radiation model in an overall flame model for the validation of either a combustion chemistry mechanism or a soot formation model against experimental data, since inaccu-rate calculations of radiation loss from the flame could lead to false conclusions on the quality of the combustion chemistry mechanism or the soot formation model.

Microgravity combustion has received renewed inter-ests since the late 1980s due to the availability of improved microgravity facilities (drop tower and manned space activities) and the needs to understand fire spread in microgravity for fire safety in spacecrafts and to advance our fundamental understanding of combustion science in general. The intrusion of buoyancy force is a greater impediment to combustion science due to the existence of large temperature and density differences. The existence of buoyancy on earth can prevent the observation of many important flame phenomena. For example, the existence of stable spherical premixed flames [8] and the extinction of counterflow low-stretched diffusion flame [1] were both observed in microgravity environment. In microgravity diffusion flames the residence time of a fluid parcel is much longer than that in their normal gravity counterparts due to the absence of buoyancy induced flow acceleration. In addi-tion, microgravity coflow diffusion flames are in general larger due to the enhanced role of diffusion in both the radial and streamwise directions and sootier as a result of prolonged residence time available for soot surface growth. Consequently, radiation heat transfer becomes significantly more important in microgravity than in normal earth gravity. Although the enhanced importance of radiation heat loss in microgravity has been recognized

[9], adequate quantitative understanding of the effects of thermal radiation transfer on laminar coflow diffusion flames is still lacking. A very recent numerical study by Kong and Liu[10]demonstrated that radiation transfer in microgravity laminar coflow methane/air diffusion flame can be so important that the flame temperature is lowered by more than 800 K due to radiation heat loss and the flame experiences partial extinction in the centerline region when the coflow air velocity is suffi-ciently low. It is noticed that the flame investigated by Kong and Liu[10]is relatively large, being about 9 cm tall at microgravity. The importance of radiation absorption in microgravity flames, however, was not investigated in the study of Kong and Liu[10].

This study attempts to gain further understanding of the importance of thermal radiation transfer and radiation absorption in a laminar coflow ethylene/air diffusion flame at both normal- and microgravity. The objectives of the present work are to quantify the effects of radiation

heat loss and radiation absorption in flame temperature and soot loading through numerical modelling. Numerical calculations of the coflow laminar ethylene/air diffusion flame were conducted using the GRI-Mech 3.0 gas-phase reaction mechanism with the removal of species and reactions related to NOxformation (except N2), the SNBCK

non-grey gas radiation model, and an acetylene based simplified soot formation model employed in previous studies[2,6,10].

2. Numerical method

The elliptic governing equations of mass, momentum, energy, and species in the low Mach number limit and in axisymmetric cylindrical coordinates (r, z) described in Refs.[11,12]are solved. The gravitational term is included in the momentum equation along z (vertically upwards) direction. The effects of buoyancy on the flame structure and soot formation characteristics under different gravity levels can be easily modelled by varying the value of the gravitational acceleration, which is 980 cm/s2 at 1g and 0 cm/s2at microgravity or 0g. The method of correction

diffusion velocity described in Ref.[13]was employed to ensure that the net diffusion flux of all species sums to zero in both r and z directions. Note that the correction velocity accounts for the thermophoretic velocity of soot. The interaction between the soot chemistry and the gas-phase chemistry was accounted for through the reaction rates of the species related to soot formation and oxidation, namely C2H2, CO, H2, O2, O, H, and OH in the

present context. Only the thermal diffusion velocities of H2 and H were taken into account using the expression

given in Ref.[13]. The source term in the energy equation due to radiation heat transfer was included and calculated using the method described below.

A modified version of the semi-empirical two-equation formulation of soot kinetics proposed by Leung et al.[14]

was used to model soot nucleation, surface growth, and oxidation. In this simplified soot nucleation and growth model, it is assumed that acetylene is the only soot nucleation and growth species, and soot nucleation and surface growth proceed, respectively, via C2H2-2C(S) + H2

and C2H2+ nC(S)-(n+2)C(S)+H2. The rates of nucleation

and surface growth are written as R1= k1(T)[C2H2] (kmol/

m3/s) and R

2=k2(T)As0.5[C2H2] (kmol/m3/s), where As=

p

(6/

p

)2/3

r

C(S)2/3 Ys2/3

r

N1/3 is the soot surface area per unit

volume and [C2H2] is the mole concentration of acetylene.

The nucleation and growth rate constants used in the present calculations were taken from a previous study [2]: k1=1000 exp( 16,103/T) [1/s] and k2=1750

exp( 10,064/T) [m0.5/s]. The density of soot

r

C(S) is

assumed to be 1.9 g/cm3. It is noted that the soot surface

growth rate is assumed to be proportional to the square root of soot surface area based on the recommendation of Leung et al.[14]. Such a sub-linear dependence of the soot surface growth rate on soot surface area can be related to the loss of active sites as soot particles mature and the agglomeration process, which reduces the available sur-face area for sursur-face growth reactions. Soot oxidation was assumed to proceed through the following reactions:

F. Liu et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 112 (2011) 1241–1249

(4)

O2+ 0.5C(S)-CO, OH+C(S)-CO+H, and O+C(S)-CO.

Further details of the governing equations and the soot model can be found in Ref.[2].

The radiation source term in the energy equation was calculated using the discrete-ordinates method (DOM) in axisymmetric cylindrical geometry. The corresponding T3

quadrature given in Ref. [6] was used for angular discretisation. Spatial discretisation of the radiative trans-fer equation was achieved using the finite volume method along with the central difference scheme. The efficient implementation approach of the statistical narrow-band correlated-k (SNBCK) based wide band model described by Liu and Smallwood [15] was employed to obtain the absorption coefficients of the combustion products con-taining CO, CO2, H2O, and soot at each wide band. In this

model, the radiative properties of CO, CO2, and H2O are

evaluated using the SNB parameters of Soufiani and Taine

[16], while the spectral absorption coefficient of soot was obtained by Rayleigh’s theory for small particles and assumed to be 5.5fv

n

, fvbeing the soot volume fraction

and

n

the wave number at the band centre.

Density of the mixture (including soot) was calculated using the ideal gas state equation. Gas-phase combustion chemistry was modelled using the GRI-Mech 3.0 mechan-ism[17], which was optimized for methane combustion, with the removal of species and reactions related to NOx formation (except N2). This simplified GRI-Mech 3.0

mechanism contains 36 species and 219 reactions. Thermal and transport properties of species and the mixture were obtained from CHEMKIN subroutines and the GRI-Mech 3.0 database. To allocate the flame front where the mixture fraction is stoichiometric, the mixture fraction field was calculated using the definition of Bilger et al.[18].

The fully coupled transport equations were discretized by the standard finite volume method in axisymmetric cylindrical coordinates. The SIMPLE algorithm for pres-sure–velocity coupling was employed. The discretized equations of momentum, pressure correction, tempera-ture, soot mass fraction, and soot number density were solved using the tri-diagonal matrix algorithm. Transport equations of gas-phase species were solved simulta-neously at each control volume using a direct solver to effectively deal with the stiffness of the detailed chem-istry to ensure convergence.

This simplified soot model, when used in conjunction with the GRI-Mech 3.0 mechanism[17], has been shown to perform quite well in the prediction of laminar coflow C2H4/air [2] and CH4/air [6] diffusion flames at normal

gravity and atmospheric pressure in terms of the visible flame height and the peak soot volume fraction against available experimental data. In addition, the effect of gravity on soot formation is a physical phenomenon, rather than a chemical one. For these reasons, it is adequate to use this simplified soot formation model for the purposes of the present study.

3. Results and discussion

The laminar ethylene/air flame investigated in this study is generated using an axisymmetric coflow diffusion

flame burner in which the fuel stream is delivered through the central fuel pipe of 10.9 mm internal diameter (0.9 mm thickness), while the coflowing air stream is delivered through the co-annular region between the fuel pipe and an outer pipe of 26 cm inner diameter. The fuel tube size considered in this study is identical to that of the G ¨ulder et al.[19]burner and very similar to that of the popular Santoro et al.[20] burner (11.1 mm id). Uniform inlet temperatures of 300 K are assumed for both the fuel and air streams. The tempera-ture at the burner rim is also kept at 300 K. The velocity distribution at the exit of the fuel pipe is assumed to be uniform at mean velocity vF. The mean fuel stream

velocity is fixed at

n

F= 3.26 cm/s throughout this study.

The inlet velocity at the air stream remains uniform at the specified velocity outside the boundary layer of the outer surface of the fuel pipe. A boundary layer type velocity distribution close to the outer surface of the fuel pipe is assumed. The air stream velocity outside the boundary layer of the outer surface of the fuel pipe is kept at vA=5 cm/s in all the calculations.

Non-uniform grids are used in both r and z directions to provide greater resolution in the large gradient regions without an excessive increase in the computing time. Fine grids are placed in the r direction between 0 and 1 cm with a grid resolution of about 0.16 mm. In the stream-wise direction, fine grids are used in the burner exit region and the grid gradually becomes coarser as the distance from the burner exit increases. The dimensions of the solution domain are 9.78 cm in the streamwise direction and 13 cm in the radial direction. In total, the computa-tional domain is divided into 251 (z)  135 (r) control volumes. Inlet conditions are specified for the fuel and air streams at the z =0 boundary. Symmetry conditions are used at the centerline, i.e. r = 0. Free slip conditions are assumed for the velocity at the outer r = 13 cm boundary. Zero-gradient conditions are enforced at the top exit boundary. In the calculation of radiation heat transfer, all the boundaries are assumed to be cold at 300 K and black.

3.1. Comparison of the flame structure and soot field at 1g and 0g

To appreciate the enhanced role of radiation heat transfer in laminar diffusion flames at microgravity, it is useful to first compare the flame structure and soot distribution of the ethylene/air diffusion flame at 1g and 0g. To highlight the variation in flame structure when the gravity level is changed from 1g to 0g, several represen-tative quantities are displayed in Figs. 1 and 2 for comparison. In both cases, the flame location, character-ized by the stoichiometric mixture fraction of fst= 0.0634,

is indicated by the black line in Figs. 1 and 2(b). Several observations can be made between the flame at 1g and 0g. First, the axial velocities at 0g (Fig. 2(a)) are much lower than those at 1g (Fig. 1(a)) due to the absence of buoyancy induced acceleration. The largest velocities at 0g occur low in the flame, as a result of flow acceleration due to heat release. The much lower axial velocities at 0g result in much longer residence times available for diffusion,

(5)

thermal radiation, and soot surface growth and oxidation. Second, the flame becomes taller and much wider at 0g, which is attributed to the enhanced role of diffusion and prolonged residence times. Finally, the flame tip is closed at 1g, but open at 0g. This is indicated by the OH mass fraction distributions shown in Figs. 1 and 2(c). Conse-quently, the flame at 1g does not emit soot, i.e. soot is completely oxidized, but emits soot at 0g. This point will be further discussed later. As a result of the emission of soot from the flame at 0g, the mass fractions of CO2at the

centerline region in the upper portion of the flame are much lower than those at 1g. The mass fractions of H2O

remain more or less unchanged when the gravity level is changed from 1g to 0g.

To illustrate how gravity affects quantities related to soot, Figs. 3 and 4 display the distributions of C2H2mass

fraction, soot volume fraction, soot nucleation rate (g cm 3s 1), and soot surface growth rate (g cm 3s 1)

at 1g and 0g, respectively. It is first noticed that the concentrations of C2H2 become much lower when the

gravity is lowered from 1g to 0g, especially in the centerline region. The distribution of soot volume fraction at 0g exhibits distinct features from that at 1g. First, soot exists only in a narrow co-annular region around the burner rim radially. This is due to the lack of radially inward flow associated with the absence of buoyancy induced flow acceleration at 0g and the lack of soot nucleation at the centerline region as a result of radiation

0 1 2 r, cm r, cm r, cm r, cm r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 3 4 5 6 7 8 z, cm u-velo, cm/s Peak: 224 cm/s 0 1 2 Temperature Peak: 1978.6K 0 1 2 CO2 Peak: 0.226 0 1 2 OH x 1.0E3 Peak:2.756E-3 0 1 2 H2O Peak: 0.086 0 1 2 3 4 5 6 7 8 z, cm

Fig. 1. Distributions of axial velocity, temperature, OH mass fraction, CO2mass fraction, and H2O mass fraction at 1g with full radiative transfer.

u-velo, cm/s Peak: 20.94 cm/s Temperature Peak: 1863.7K CO2 Peak: 0.194 OH x 1.0E3 Peak:1.952E-3 H2O Peak: 0.086 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm

Fig. 2. Distributions of axial velocity, temperature, OH mass fraction, CO2mass fraction, and H2O mass fraction at 0g with full radiative transfer. F. Liu et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 112 (2011) 1241–1249

(6)

heat loss. Second, the soot volume fractions are much higher at 0g (with a peak value more than 90 ppm vs. a peak value of 8.85 ppm at 1g) and soot escapes from the flame,Fig. 4(b), as mentioned above, instead of being fully oxidized with the flame at 1g,Fig. 3(b). It is interesting to observe that both the soot nucleation and surface growth rates at 0g are actually lower than those at 1g, compare

Fig. 4(c) and (d) with Fig. 3(c) and (d). It is therefore evident that the enhanced soot loading at 0g can be attributed to the combined effect of prolonged residence times, as a result of much lower axial velocity shown in

Fig. 2(a), and the absence of soot oxidation by OH. The latter point can be appreciated by noticing that the region of high OH concentration does not intersect or overlap with the path along which soot is transported. On the other hand, the high OH concentration region fully envelops the sooting region with a closed tip at 1g. This is why the flame at 1g does not emit soot.

3.2. Effect of radiation heat loss

To quantify the importance of radiation heat loss in the laminar coflow ethylene/air diffusion flame investigated in this study, calculations at both 1g and 0g were also conducted with the radiative source term removed from the energy equation, i.e. without radiation heat loss. A quantitative understanding of the importance of radiation heat transfer in this flame at 1g and 0g can then be gained by comparing the results with and without radiation heat transfer. Figs. 5 and 6 display the distributions of axial velocity, temperature, OH mass fraction, CO2 mass

fraction, and H2O mass fraction at 1g and 0g, respectively.

Similarly, Figs. 7 and 8 show the distributions of C2H2

mass fraction, soot volume fraction, soot nucleation rate, and soot surface growth rate. Although radiation heat loss has significant impact on the flame structure and soot field in this flame at 1g, compare Figs. 5 and 7 with

R1 x 1.0E7 Peak: 4.817E-7 R2 x 1.0E4 Peak: 6.974E-4 C2H2 Peak: 0.0576

Soot vol. frac. Peak: 8.852 ppm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm

Fig. 3. Distributions of C2H2mass fraction, soot volume fraction, soot nucleation rate (g cm 3s 1), and soot surface growth rate (g cm 3s 1) at 1g with

full radiative transfer.

R1 x 1.0E7 Peak: 2.512E-7 R2 x 1.0E4 Peak: 3.262E-4 C2H2 Peak: 0.0407

Soot vol. frac. Peak: 92.2 ppm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm

Fig. 4. Distributions of C2H2mass fraction, soot volume fraction, soot nucleation rate (g cm 3s 1), and soot surface growth rate (g cm 3s 1) at 0g with

(7)

Figs. 1 and 3, it has a much stronger influence at 0g. At 1g, radiative heat loss lowers the flame temperature by about 144 K and increases the visible flame height from about 4.5 to about 6 cm. At 0g, radiation heat loss reduces the peak flame temperature by as much as 240 K, compare

Fig. 6(b) with Fig. 2(b), and the flame changes from a close-tipped one to an open-tipped one, and from a non-smoking one to a non-smoking (soot-emitting) one, compare

Fig. 8(b) to Fig. 4(b). In both cases (1 and 0g), radiation heat loss lowers the temperatures in the centerline region the most, especially at 0g, where the flame temperatures in the upper portion of the flame around z= 6 cm is lowered by more than 1000 K due to radiation heat loss. The corresponding temperature drop at 1g is about 300 K. The much stronger effect of radiation heat loss

at 0g than that at 1g has also been recently shown by Kong and Liu[10].

3.3. Effect of radiation absorption

The importance of radiation absorption was also quantified by conducting additional calculations with the absorption term in the radiative source term removed, i.e. emission only. This is equivalent to the optically thin approximation. Due to space limitation, results for the axial velocity, temperature, and mass fractions of OH, CO2,

and H2O are not shown. Overall the effect of radiation

absorption was found to be relatively small. It alters the flame structure only quantitatively at both 1 and 0g, but u-velo, cm/s Peak: 241.07 cm/s Temperature Peak: 2122.4 K OH x 1.0E3 Peak: 3.066 CO2 Peak: 0.214 H2O Peak: 0.0894 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm

Fig. 5. Distributions of axial velocity, temperature, OH mass fraction, CO2mass fraction, and H2O mass fraction at 1g without radiative transfer.

U 20 19.2 18.4 17.6 16.8 16 15.2 14.4 13.6 12.8 12 11.2 10.4 9.6 8.8 8 7.2 6.4 5.6 4.8 4 3.2 2.4 1.6 0.8 0 OH 2 1.92 1.84 1.76 1.68 1.6 1.52 1.44 1.36 1.28 1.2 1.12 1.04 0.96 0.88 0.8 0.72 0.64 0.56 0.48 0.4 0.32 0.24 0.16 0.08 0 CO2 0.225 0.216 0.207 0.198 0.189 0.18 0.171 0.162 0.153 0.144 0.135 0.126 0.117 0.108 0.099 0.09 0.081 0.072 0.063 0.054 0.045 0.036 0.027 0.018 0.009 0 H2O 0.085 0.0816 0.0782 0.0748 0.0714 0.068 0.0646 0.0612 0.0578 0.0544 0.051 0.0476 0.0442 0.0408 0.0374 0.034 0.0306 0.0272 0.0238 0.0204 0.017 0.0136 0.0102 0.0068 0.0034 0 2000 1932 1864 1796 1728 1660 1592 1524 1456 1388 1320 1252 1184 1116 1048 980 912 844 776 708 640 572 504 436 368 300 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm u-velo, cm/s Peak: 21.37 cm/s Temperature Peak: 2103.2K CO2 Peak: 0.229 OH x 1.0E3 Peak:2.488 H2O Peak: 0.102

Fig. 6. Distributions of axial velocity, temperature, OH mass fraction, CO2mass fraction, and H2O mass fraction at 0g without radiative transfer. F. Liu et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 112 (2011) 1241–1249

(8)

not qualitatively as the radiation heat loss does at 0g. The effect of radiation absorption at 1g is similar to that shown in a previous study[2]. At 0g, however, the effect of absorption is enhanced as indicated by the following: (i) the peak flame temperature is further lowered from 1863.7 to 1851.4 K, (ii) the flame height indicated by the location of the stoichiometric mixture fraction on the flame centerline is further increased from 7.2 to 7.8 cm, and (iii) the temperatures in the centerline region in the upper portion of the flame of around z =7 cm are further lowered by about 100 K. Results for C2H2mass fraction,

soot volume fraction, soot nucleation rate, and soot surface growth rate at 1 and 0g are displayed in Figs. 9 and 10, respectively. The effects of radiation absorption on these quantities at 1 and 0g can be observed by comparing

Fig. 9withFig. 3andFig. 10withFig. 4, respectively. At 1g the effects of absorption on these quantities are fairly

small, consistent with the above discussion for the other five quantities. At 0g, however, radiation absorption affects the flame height and the soot loading significantly. The peak soot volume fraction is reduced from 92 to 76.7 ppm when radiation absorption is neglected.

4. Conclusions

Detailed numerical modelling of a laminar coflow ethylene/air diffusion flame was conducted at both normal and microgravity and atmospheric pressure with three radiation transfer treatments, namely the full radiative transfer, no radiation, and no radiation absorp-tion. The effects of radiation heat loss were found to be significantly more important at microgravity than at normal gravity. For the flame investigated in this study,

0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm

Peak: 7.091E-7 Peak: 13.751E-4 Peak: 0.0753

Soot vol. frac.

Peak: 9.115 ppm R1 x 1.0E7 R2 x 1.0E4 C2H2

Fig. 7. Distributions of C2H2mass fraction, soot volume fraction, soot nucleation rate (g cm 3s 1), and soot surface growth rate (g cm 3s 1) at 1g

without radiative transfer.

0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm

Peak: 3.935E-7 Peak: 12.154E-4 Peak: 0.0608

Soot vol. frac.

Peak: 28.6 ppm R1 x 1.0E7 R2 x 1.0E4 C2H2

Fig. 8. Distributions of C2H2mass fraction, soot volume fraction, soot nucleation rate (g cm 3s 1), and soot surface growth rate (g cm 3s 1) at 0g

(9)

radiation heat loss completely alters the flame structure at microgravity from a tip-closed flame to a tip-open one and from a non-smoking flame to a smoking one. Although radiation heat loss was also found important at normal gravity, especially in the upper portion of the flame, its effect is much smaller compared to that at microgravity. Radiation absorption was also found to be more important at microgravity than at normal gravity. This study once again demonstrates that it is essential to model radiative heat transfer accurately in laminar diffusion flame modelling at microgravity.

Acknowledgement

Wenjun Kong would like to acknowledge the financial support from National Natural Science Foundation of China (NSFC) under Grant no. 50936005.

References

[1] Maruta K, Yoshida M, Guo H, Ju Y, Niioka T. Extinction of low-stretched diffusion flame in microgravity. Combust Flame 1998;112:181–7.

[2] Liu F, Guo H, Smallwood GJ, G ¨ulder O¨ L. Effects of gas and soot radiation on soot formation in a coflow laminar ethylene diffusion flame. J Quant Spectrosc Rad Trans 2002;73:409–21.

[3] Liu F, Guo H, Smallwood GJ, G ¨ulder O¨ L. Effects of gas and soot radiation on soot formation in counterflow ethylene diffusion flames. J Quant Spectrosc Rad Trans 2004;84:501–11.

[4] Kaplan CR, Baek SW, Oran ES, Ellzey JL. Dynamics of a strongly radiating unsteady ethylene jet diffusion flame. Combust Flame 1997;96:256–63.

[5] Sivathanu YR, Gore JP. Coupled radiation and soot kinetics calculations in laminar acetylene/air diffusion flames. Combust Flame 1994;97:1–21.

[6] Liu F, Guo. H, Smallwood GJ. Effects of radiation model on the modeling of a laminar coflow methane/air diffusion flame. Combust Flame 2004;138:136–54.

[7] Wang L, Haworth DC, Turns SR, Modest MF. Interactions among soot, thermal radiation, and NOx emissions in oxygen-enriched

0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm

Peak: 4.729E-7 Peak: 6.703E-4 Peak: 0.0564

Soot vol. frac. Peak: 8.724 ppm

R1 x 1.0E7 R2 x 1.0E4

C2H2

Fig. 9. Distributions of C2H2mass fraction, soot volume fraction, soot nucleation rate (g cm 3s 1), and soot surface growth rate (g cm 3s 1) at 1g

without radiative absorption.

0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm 0 1 2 r, cm 0 1 2 3 4 5 6 7 8 z, cm

Peak: 2.399E-7 Peak: 3.008E-4 Peak: 0.0399

Soot vol. frac.

Peak: 76.69 ppm R1 x 1.0E7 R2 x 1.0E4 C2H2

Fig. 10. Distributions of C2H2mass fraction, soot volume fraction, soot nucleation rate (g cm 3s 1), and soot surface growth rate (g cm 3s 1) at 0g

without radiative absorption.

F. Liu et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 112 (2011) 1241–1249

(10)

turbulent nonpremixed flames: a computational fluid dynamics modeling study. Combust Flame 2005;141:170–9.

[8] Ronney PD. Near-limit flame structures at low lewis number. Combust Flame 1990;82:1–14.

[9] Shu Z, Choi CW, Aggarwal SK, Katta VR, Puri IK. Gravity effects on steady two-dimensional partially premixed methane-air flames. Combust Flame 1999;118:91–107.

[10] Kong W, Liu, F. Numerical study of the effects of gravity on soot formation on lamminar coflow methane/air diffusion flames under different air stream velocities. Combust Theory Modelling 2009;13(6):993–1023.

[11] Liu F, Guo H, Smallwood GJ, G ¨ulder O¨ L. Numerical moddelling of soot formation and oxidation in laminar coflow non-smoking and smoking ethylene diffusion flames. Combust Theory Modelling 2003;7:301–15.

[12] Guo H, Liu F, Smallwood GJ, G ¨ulder O¨ L. The flame preheating effect on numerical modelling of soot formation in a two-dimensional laminar ethylene–air diffusion flame. Combust Theory Modelling 2002;6:173–87.

[13] Kee RJ, Grcar JF, Smooke MD, Miller JA. A Fortran program for modeling steady laminar one-dimensional premixed flames. Report no. SAND85-8240, Sandia National Laboratories, 1985.

[14] Leung KM, Lindstedt RP, Jones WP. A simplified reaction mechan-ism for soot formation in nonpremixed flames. Combust Flame 1991;87:289–305.

[15] Liu F, Smallwood GJ. An efficient approach for the implementation of the SNB based correlated-k method and its evaluation. J Quant Spec Rad Trans 2004;84:465–75.

[16] Soufiani A, Taine J. High temperature gas radiative property parameters of statistical narrow-band model for H2O, CO2 and

CO, and correlated-K model for H2O and CO2. Int J Heat Mass

Transfer 1997;40:987–91.

[17] Smith GP, Golden DM, Frenklach M, Moriarty NW, Eiteneer B, Goldenberg M, Bowman CT, Hanson RK, Song S, Gardiner Jr, WC, Lissianski VV, Qin Z., /http://www.me.berkeley.edu/gri_mech/S. [18] Bilger RW, St ˚arner SH, Kee RJ. On reduced mechanisms for

methane–air combustion in nonpremixed flames. Combust Flame 1990;80:135–49.

[19] G ¨ulder O¨ L, Snelling DR, Sawchuk RA. Influence of hydrogen addition to fuel on temperature field and soot formation in diffusion flames. Proc Combust Inst 1996;26:2351–8.

[20] Santoro RJ, Yeh TT, Horvath JJ, Semerjian HG. The transport and growth of soot particles in laminar diffusion flames. Combust Sci Technol 1987;53:89–115.

Figure

Fig. 1. Distributions of axial velocity, temperature, OH mass fraction, CO 2 mass fraction, and H 2 O mass fraction at 1g with full radiative transfer.
Fig. 3. Distributions of C 2 H 2 mass fraction, soot volume fraction, soot nucleation rate (g cm 3 s 1 ), and soot surface growth rate (g cm 3 s 1 ) at 1g with full radiative transfer.
Fig. 5. Distributions of axial velocity, temperature, OH mass fraction, CO 2 mass fraction, and H 2 O mass fraction at 1g without radiative transfer.
Fig. 8. Distributions of C 2 H 2 mass fraction, soot volume fraction, soot nucleation rate (g cm 3 s 1 ), and soot surface growth rate (g cm 3 s 1 ) at 0g without radiative transfer.
+2

Références

Documents relatifs

Distribution automatique Distribution automatique centralisée centralisée Firewall 1 Firewall 3 Officier de sécurité Politique de contrôle d ’accès Modèle du

Even though almost all of these “Big YouTubers” have started creating videos when they were teenagers themselves, the platform has changed significantly in the

Supervised models tend to perform well on identifying specific signals comparing to a consistent background, and is probably inappropriate for anomaly detection with

Our con- tribution in this paper is three-fold: Firstly, we propose a new framework (Section 3) that is capable of object detection by fea- ture aggregation from large point

The W +jets MC samples are used in the analysis according to the number of heavy flavor (c or b) jets in the final state, classified as follows: W + light de- notes events without c

Deux ajustements de ce gradient découlent du comportement des consommateurs. 69) rapporte que pour certains enquêtés, des fruits et légumes sont considérés comme locaux

En France, les chevaux sont présents sur tout l’hexagone, et plus particuliè- rement dans les zones herbagères du Grand-Ouest (notamment en Basse- Normandie) et de moyennes

We also con- sider a Bonded Particle Model (BPM) in which each par- ticle is represented by an aggregate of primary rigid parti- cles with appropriate long-range interactions such