• Aucun résultat trouvé

Structural response of different Lewis number premixed flames interacting with a toroidal vortex

N/A
N/A
Protected

Academic year: 2021

Partager "Structural response of different Lewis number premixed flames interacting with a toroidal vortex"

Copied!
9
0
0

Texte intégral

(1)

HAL Id: hal-01993705

https://hal.archives-ouvertes.fr/hal-01993705

Submitted on 10 Mar 2019

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

Structural response of different Lewis number premixed

flames interacting with a toroidal vortex

Chaimae Bariki, Fabien Halter, Fabien Thiesset, Christian Chauveau,

Iskender Gökalp, Ketana Teav, Sina Kheirkhah, Adam Steinberg

To cite this version:

Chaimae Bariki, Fabien Halter, Fabien Thiesset, Christian Chauveau, Iskender Gökalp, et al..

Struc-tural response of different Lewis number premixed flames interacting with a toroidal vortex.

Proceed-ings of the Combustion Institute, Elsevier, 2019, 37 (2), pp.1911-1918. �10.1016/j.proci.2018.06.114�.

�hal-01993705�

(2)

Structural response of di

fferent Lewis number premixed flames interacting with a

toroidal vortex

Chaimae Barikia, Fabien Haltera,∗, Fabien Thiesseta, Christian Chauveaua, Iskender Gökalpa, Ketana Teavb, Sina Kheirkhahc, Adam M. Steinbergb

aICARE-CNRS, Avenue de la Recherche Scientifique, Orléans Cedex2 45072, France

bInstitute for Aerospace Studies, University of Toronto, 4925 Dufferin Street, Toronto, Ontario M3H 5T6, Canada cSchool of Engineering, The University of British Columbia, 1137 Alumni Avenue, Kelowna, British Columbia, V1V1V7, Canada

Abstract

Simultaneous measurements of temperature, CH* and OH* chemiluminescent species are carried out to explore the impact of stretch rate and curvature on the structure of premixed flames. The configuration of an initially flat premixed flame interacting with a toroidal vortex is selected for the present study and reasons for this choice are discussed. Lewis number effects are assessed by comparing methane and propane flames. It is emphasized that the flame structure experiences very strong variations. In particular, the flame is shrunk (broadened) in the initial (final) period of the interaction with the vortex where strain rate (curvature) contribution of the stretch rate is predominant. By further analysing independently the thickness of the preheat and reaction zones, it is shown that for propane flames, not only the former but also the latter is significantly altered in zones where the flame curvature is negative. Changes in the reaction zone properties are further emphasized using CH* and OH* radicals. It is demonstrated that higher thermal diffusivity plays a significant role around curved regions, in which the enhanced diffusion of heat leads to a strong increase of CH* compared to OH* intensity. As an overall conclusion, this study suggests that it would be interesting to reassess the internal flame structure at lower and moderate Karlovitz numbers since changes might appear for a moderate vortex intensity with typical size much larger than the flame thickness.

Keywords:

Flame-vortex interaction, Laminar premixed flames, Curvature/stretch effects, Flame front thinning/thickening, Chemiluminescence

Corresponding author:

Email address: fabien.halter@cnrs-orleans.fr (Fabien Halter)

(3)

1. Introduction

1

When a premixed flame evolves in a non-uniform

2

flow field, e.g. a turbulent flame, its geometry

3

(flame surface area, curvature), kinematics (flame

dis-4

placement and consumption speed), structure

(ther-5

mal, preheat, and reaction zone thicknesses) and

6

thermo-chemistry (species and temperature

trans-7

port/production) can be altered. These flame features

8

are coupled in a way that remains only partially

under-9

stood, and hence a closed set of equations that

accu-10

rately describe premixed flame behavior remains a

re-11

search objective. Notwithstanding the complex nature

12

of the fully turbulent problem, simplified - though

in-13

sightful - configurations can yield insight regarding the

14

elementary, leading order mechanisms at play in real

15

turbulent premixed flames. From this perspective, the

16

configuration of a single vortex interacting with an

ini-17

tially flat premixed flame is well-suited. In our opinion,

18

this flow situation is equally relevant for fundamental

19

purposes as steady laminar stretched flames or turbulent

20

flames embedded in homogeneous isotropic turbulence.

21

Indeed, the literature demonstrates that flame/vortex

22

interactions (FVI) have significantly helped in better

un-23

derstanding the effect of vortex strength on flame stretch

24

and quenching [1, 2]. Such studies have led to the

de-25

velopment of the so-called efficiency functions [3–6],

26

which remain widely employed in Large Eddy

Simu-27

lations (LES) of turbulent premixed combustion. Time

28

transient effects [7–9] have also been evidenced using

29

FVIs. More recently, Thiesset et al. [10] provided the

30

first experimental confirmation of the asymptotic theory

31

of Refs. [11, 12], revealing that flame speeds are

un-32

equally altered by flame curvature and hydrodynamic

33

straining. Hence, two Markstein numbers should be

34

used, one characterizing the dependence of flame speed

35

to strain, the other to curvature. This demonstrates the

36

utility of simultaneously observing strain and curvature

37

in the same system; such results could not be obtained

38

in, e.g, spherically expanding flames or steady

stagna-39

tion point flames.

40

However, one aspect of FVI that has received less

at-41

tention is the effect of vortex-induced strain and

curva-42

ture on the flame structure and thermo-chemistry. The

43

present study aims to fill this gap by experimentally

44

determining the thermal structure and some indicators

45

of diffusion and chemistry during FVI. The general

46

goal is to use FVI to emphasize the effect of

differen-47

tial diffusion, i.e. the imbalance between thermal and

48

molecular diffusivities as defined by the Lewis number

49

(Le= α/Dd, where α is the thermal diffusivity and Dd

50

is the molecular diffusivity of the limiting reactant), in a

51

curved and strained flame.

52

Previous fundamental studies on differential diffusion

53

effects have focused on e.g. the tip of a Bunsen flame

54

[13–16]. This zone provides a convenient

configura-55

tion to examine distortions between the temperature and

56

species path-lines. Law et al. [13] and Mizomoto et al.

57

[14] performed experiments using different fuels (e.g.

58

methane, propane, and hydrogen) to demonstrate that

59

negatively curved (concave towards the reactants)

non-60

unity Lewis number flames exhibit changes not only in

61

the internal flame structure, but also the flame burning

62

rate. This result was later supported by DNS studies of

63

Haworth and Poinsot [17] and Rutland and Trouvé [18],

64

which considered a freely propagating premixed flame

65

in homogeneous isotropic turbulence using single-step

66

Arrhenius kinetics chemistry. They showed that

nega-67

tively curved flame elements tend to have lower local

68

flame speeds for Le < 1, whereas the correlation is

op-69

posite for Le > 1. Fewer studies have directly examined

70

the internal flame structure of such flames. In DNS of

71

equidiffusive flames (i.e. Le = 1), Poinsot et al. [15]

72

found that the preheat zone thickness increases in

re-73

gions of negative curvature, while the reaction zone

re-74

mains unaffected. However, there remains a knowledge

75

gap regarding the internal flame structure for Le , 1.

76

The present study focuses on understanding the flame

77

thermal structure and chemistry by carrying out

simul-78

taneous laser Rayleigh scattering thermometry, OH*

79

chemiluminescence, and CH* chemiluminescence

mea-80

surements in a FVI configuration. The impact of di

ffer-81

ential diffusion is assessed by comparing a

stoichiomet-82

ric methane flame (Le ≈ 1) and a slightly lean propane

83

flame (Le > 1), while keeping the vortex

characteris-84

tics relative to the flame constant. That is the ratio

be-85

tween the vortex rotational velocity and laminar flames,

86

as well as the ratio between vortex size and flame

thick-87

ness are held nearly constant in this study. Note that the

88

case of Le < 1 is not investigated here because these

89

flames exhibit cellular thermodiffusive instabilities that

90

alter the initial conditions for the FVI.

91

2. Experimental set-up

92

2.1. Apparatus and diagnostic configuration

93

Experiments were performed at the University of

94

Toronto, using the FVI burner from Refs. [6, 10]. A

95

flat premixed flame is stabilized in a single jet

stagna-96

tion flow configuration before interacting with a toroidal

97

vortex of prescribed strength. Premixed fuel and air are

98

first laminarized using a grid to prevent turbulent

struc-99

tures from being convected in the burner. The reactive

100

(4)

mixture then passes through a converging section

hav-101

ing an exit diameter of 15 mm. The velocity profile at

102

the exit of the contraction is nearly top hat. A burner

103

to stagnation plate distance of 25 mm was chosen in

or-104

der to minimize the tangential strain rate before the FVI,

105

while ensuring that the flame is repeatably stabilized. A

106

laminar coaxial co-flow of nitrogen is used to avoid the

107

influence of external perturbations.

108

The toroidal vortex generator consists of a 2 mm tube

109

located on the centerline and ending 35 mm upstream

110

the burner exit. The vortex is generated by applying

111

a sudden pressure discharge of reactive mixture at the

112

same equivalence ratio as the main flow. For this

pur-113

pose, the tube from which the vortex emanates is

con-114

nected to a pressurized tank via an electrovalve, which

115

enables to control the vortex strength. Caution was

116

taken to ensure that the vortex had identical

composi-117

tion as the jet flow. This is done to avoid any

discon-118

tinuities or chemiluminescence signal variations due to

119

inhomogeneous equivalence ratio [19].

120

Measurements consisted of simultaneous 2D laser

121

Rayleigh scattering (LRS) thermometry and

chemilu-122

minescence (CL) measurements of CH* and OH*. The

123

setup is illustrated in Fig. 1.

124

Figure 1: Optical configuration. CL: cylindrical lens, DMx: dichroic

mirror, SLx: spherical lens, Fx: bandpass filter.

The LRS and CH* CL systems operated at a 10 Hz

125

repetition rate. In contrast, the vortex passed through

126

the flame in a few milliseconds. Hence, a single LRS

127

and CH* measurement was obtained during each vortex

128

passage. It previously has been shown that the

exper-129

iment is highly repeatable [6, 10], and hence the

en-130

tire time sequence of the interaction can be obtained

131

by compiling independent measurements at different

de-132

lays (phases) after the vortex injection. Measurements

133

were obtained with 0.3 ms intervals, resulting in at least

134

21 phases (depending on conditions). Five iterations

135

were performed at each phase to confirm

repeatabil-136

ity and compile phase-resolved statistics. Note that the

137

OH* CL images were obtained at 10 kHz repetition rate,

138

and hence provide the complete interaction sequence for

139

each vortex passage.

140

The LRS systems consisted of a 532 nm Nd:YAG

141

laser with a typical energy of 430 mJ/pulse (Spectra

142

Physics Lab 170) and a sCMOS camera (Andor Zyla),

143

operating at full resolution (2160 × 2560 pixels2) and

144

with a 10µs exposure. The camera was fitted with a

145

commercial lens (Tamron, f /#= 3.5, f = 180 mm) and

146

a 532±10 nm interference filter (F1). In order to account

147

for shot-to-shot laser power and profile variations, a

ref-148

erence sCMOS camera (Andor Neo) was used to record

149

a jet of clean air simultaneously with each burner LRS

150

image. The laser sheet was formed using two spherical

151

lenses ( fS L1 = −100 mm, fS L2 = 150 mm) and a cylin-152

drical lens ( fCL = 750 mm), resulting in a beam waist

153

of 80µm and a height of 12 mm.

154

The flame luminosity is split by a dichroic mirror

155

(DM1), which reflects the ultraviolet light from OH*

156

and transmits the visible light from CH*. The OH* CL

157

system consists of a high-speed camera (Photron SA-Z,

158

10 kHz, 1024 × 1024 pixels2), image intensifier

(Invisi-159

ble Vision UVi, gate 90µs), UV lens (Nikon, f /# = 4.5,

160

f = 105 mm), and bandpass filter (F2, center

wave-161

length 310 nm, bandwidth 10 nm). The CH* system

162

consists of a sCMOS camera (Andor Zyla, 2160 × 2560

163

pixels2), intensifier (LaVision IRO, gate 90µs), camera

164

lens (Tamron, f /# = 3.5, f = 180 mm), and

band-165

pass filter (F3, center wavelength 430 nm, bandwidth

166

10 nm). The cameras images were aligned using a dot

167

target. The projected pixel resolutions are 7.7, 16.7 and

168

43.3 µm/px for Rayleigh, CH*, and OH*

chemilumi-169

nescence techniques respectively. Note that all the

cam-170

eras were corrected for their corresponding whitefield

171

responses and backgrounds.

172

The chemiluminescence measurement technique

pro-173

vides a line-of-sight integrated field. Since the

configu-174

ration is highly symmetric, Abel deconvolution was

ap-175

plied to obtain planar images of CH* and OH* at the

176

mid-plane. The BAsis Set Expansion (BASEX) method

177

was used [20]. Negligible sensitivity was found to the

178

input or parameters of the BASEX algorithm.

179

2.2. Selected experimental conditions

180

Stoichiometric (φ = 1) methane/air and lean (φ =

181

0.9) propane/air mixtures were selected to study the

182

flame response to curvature and thermo-diffusive

ef-183

fects. The vortex parameters, i.e. the rotational

veloc-184

ity Uθ = 0.9 m/s, convection velocity Uc = 0.35 m/s,

185

and core-to-core distance Rv= 4 mm, were determined

186

just before the vortex starts interacting with the flame

(5)

Table 1: Experimental conditions and key parameters: adiabatic flame temperature (Tp), vortex core-to-core distance (Rv), thermal flame

thickness (δ0

L), vortex rotational velocity (Uθ), laminar flame speed

(S0

L), and effective Lewis number Lee f f.

Case Fuel φ Tp Rv/δ0L Uθ/S0L Lee f f

M10 CH4 1.0 2200 9.6 2.5 1.05

P09 C3H8 0.9 2157 11.3 2.45 1.5

using previous measurements [6, 10]. Flame

character-188

istics are calculated for a 1D stagnation laminar flame

189

configuration, performed using the PREMIX module in

190

CHEMKIN-II. The GRI-Mech 3.0 [21] chemical

mech-191

anism was used for the methane flame and

ARAMCO-192

Mech 2.0 [22] for the propane flames. The flame

ther-193

mal thickness is defined as δ0

L = |∇c| −1

max, where the

194

progress variable c= (T − Tr)/(Tp− Tr), Tris the

reac-195

tant temperature, and Tpis the adiabatic flame

temper-196

ature. Tpwas found to be nearly the same for the two

197

selected cases meaning that differential thermal

expan-198

sion is negligible.

199

The effective Lewis numbers (Lee f f) are evaluated

200

using the methodology of Bechtold et al. [11].

Stoi-201

chiometric methane/air mixture is thermo-diffusionally

202

neutral (i.e. Lee f f ≈ 1), whilst the lean propane flame

203

is thermo-diffusionally stable (Lee f f ≈ 1.5). Note that

204

these cases have nearly identical Uθ/SL, meaning that

205

differential flow effects are eliminated and the Lewis

206

number effect can be isolated. Table 1 outlines the key

207

parameters of the two selected conditions.

208

2.3. Temperature inversion methodology

209

Instantaneous LRS images were first binned into 4×4

210

windows in order to increase the signal-to-noise ratio,

211

and then filtered using a 8 × 8 Wiener filter. The size of

212

the filter was chosen in a way that it does not modify the

213

flame front structure.

214

Assuming isobaric flow of an ideal gas, LRS signal can be inverted to temperature via

T = Tre f

Ire f − IBG

If lame− IBG

(∂σ/∂Ω)mix

(∂σ/∂Ω)mix,re f (1) where If lame and Ire f represent the Rayleigh scattered

light from the flame and a calibration image of fresh gases at a known temperature, respectively. IBGis the

background intensity, which consists of the dark field of the camera and the laboratory background noise. The differential Rayleigh scattering cross-section (∂σ/∂Ω) of the local gas is the mole-fraction weighted sum of

the differential cross-sections of the individual species. Individual molecular cross-sections were calculated as

σi= 4π2 λ4 n − 1 N0 !2 3 3 − 4ρp !  1+ ρp  (2)

where λ is the laser wavelength, n is the index of

refrac-215

tion, N0 is the Loschmidt number and ρp is the

depo-216

larization ratio [23–25]. The variation of composition

217

with temperature was assumed to follow that of a planar

218

laminar flame, computed as described above. The

tem-219

perature uncertainty was calculated based on the

Pois-220

son noise since it is dominated by the shot-noise, and

221

is equal to 5 K and 50 K in the reactants and products,

222

respectively.

223

The resultant temperature fields were converted to

224

progress variable fields. These were used to calculate

225

the flame thermal thickness from the progress variable

226

gradient.

227

3. Results and Discussion

228

3.1. Flame interacting with a toroidal vortex

229

A time sequence of temperature fields from the FVI

230

in the propane/air flame is shown in Fig. 2. Two

some-231

what distinct time periods are apparent, one in which

232

the vortex modifies the flame and one in which the flame

233

returns to its unperturbed state. At t = 0 ms, the flame

234

was almost flat, with a slight positive curvature due to

235

the weak wake from the vortex-generating tube. For

236

0 < t . 4.5 ms, the toroidal vortex was injected,

con-237

vected to the flame, and distorted the flame front. The

238

vortex did not cause local extinction, and the flame

re-239

turned to its original shape over t& 4.5 ms. This second

240

time period is characterized by very high negative

cur-241

vatures and is driven by the consumption velocity.

242

The normalized rate-of-change in flame surface area,

243

A, constitutes the total stretch rate. The stretch-rate can

244

be decomposed into two contributions, one due to

tan-245

gential strain and one to the propagation of curved flame

246

elements. The time evolution of flame area is displayed

247

in Fig. 3 for both cases. The flame response to the FVI

248

was almost identical during the initial phase of the FVI,

249

until t ≈ 3.8 ms, which is dominated by vortex-induced

250

strain causing increased A. However, significant

dif-251

ferences occurred between the Lee f f ≈ 1 methane/air

252

flame and Lee f f > 1 propane/air flame during the

sec-253

ond phase of the FVI, in which the highly curved flame

254

returned to its original configuration.

255

(6)

400 600 800 1000 1200 1400 1600 1800

Figure 2: Time sequence of temperature field for propane flame at φ = 0.9. ∆ is the width domain. The centerline at 6.7 ms is where the axial profiles will be extracted.

0 1 2 3 4 5 6 7 0 5 10 15 20 25 30 35

Figure 3: Time evolution of flame area A(t) for P09 (red diamonds) and M10 (black circles) cases. The selected phases to be compared are circled, characterized by approximately the same curvature at the flame tip.

3.2. Flame front thinning/ thickening

256

To investigate this further, the total flame curvature,

257

κm, was calculated as the average of the in-plane

curva-258

ture κ1= r0h00− r00h0and the contribution in the

circum-259

ferential direction κ2 = h0/r. Here, r and h are the

ra-260

dial and the streamwise spatial flame front coordinates,

261

respectively, indexed against the curvilinear abscissa s,

262

and (·)0is a derivative with respect to s. Curvature was

263

extracted along the c= 0.7 contour, which is

represen-264

tative of the reaction zone.

265

To demonstrate the effects of curvature and

differen-266

tial diffusion on the inner flame structure, four phases

267

were selected from each flame for which the minimum

268

radius of curvature (i.e. along the burner centerline) was

269

nearly identical between the fuels. These phases are

cir-270

cled in Fig. 3. The first two phases correspond to the

271

early stage of FVI where strain rate is predominant over

272

curvature, whereas the two last phases correspond to

273

the second part of FVI where the flame tip is extremely

274

curved. During the early stage, the equal curvatures

oc-275

curred at the same time in the vortex passage.

How-276

ever, the curvature of the propane flame lagged that of

277

the methane flame by approximately 0.5 ms during the

278

later stage of the FVI. This lag could be explained by

279

the initial state of propane flame that is less positively

280

curved than methane flame. Recall that at steady state,

281

the flames are slightly convex towards the unburned

re-282

actants due to the vortex generator that creates a small

283

velocity deficit in its wake [10].

284

The inner flame structure during the FVI is examined

285

in Fig. 4 by the variation of |∇c| with c at the flame

cen-286

terline (r = 0) for these phases. This parameter is

in-287

versely related to the thermal flame thickness. Note that

288

for each fuel, each phase has been normalized by the

289

corresponding maximum value of |∇c| when the flame

290

is flat, i.e. at Phase 1 before the interaction takes place.

291

The first phase illustrates the initial inner flame structure

292

affected only by the relatively weak straining from the

293

stagnation flow. The curves for Phase 1 are relatively

294

symmetric around c= 0.5.

295

Phase 2 occurred during the upward movement of

296

the flame as it was strained and wrinkled by the

vor-297

tex. At this phase, |∇c| exhibits greater values compared

298

to the initial phase, indicating a relative thinning of 30

299

and 40 % for methane and propane flames, respectively.

300

In Ref. [26], it was demonstrated that equidiffusive

301

flame structure remains insensitive to strain rate using a

302

counterflow configuration. On the other hand, Driscoll

303

[27] states in his review that a positively strained lean

304

methane flame becomes thinner, whereas a lean propane

305

flame becomes thicker. The result presented here for

306

FVI brings new insights.

307

For a highly negatively curved shape at Phase 3, a

308

prominent decrease of |∇c| is identified, meaning that

309

the flame front is broadened. Furthermore, the profile

310

becomes asymmetric, with elevated |∇c| for 0.45 < c <

311

1 in the methane flame and a similar pattern, though

312

with lower amplitude, in the propane flame. When the

313

curvature decreases (Phase 4), the thickening

dimin-314

ishes for both fuels but faster for methane.

315

Changes to the internal flame structure during FVI are

316

further elucidated by decomposing the flame front into

317

two regions, one approximately describing the thickness

318

of the preheat zone and the other the reaction zone

thick-319

ness. The preheat zone thickness along the length of the

320

flame (δpre) is characterized by the minimum distance

321

between the c = 0.1 and c = 0.7 iso-contours at the

322

same phases used in Fig. 4. In Fig. 5(a), this

thick-323

ness is plotted versus the local flame curvature on the

(7)

0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1 1.2 1.4

Figure 4: Normalized |∇c| profiles versus progress variable c at the flame centerline for different phases during FVI for P09 case (filled symbols) and M10 case (open symbols).

c = 0.7 iso-contour. For Phases 1 and 2, normalized

325

curvatures are in the range −0.3. κm·δL. 0.2 and the

326

preheat zone thickness remains nearly constant.

How-327

ever, significant changes are observed for Phases 3 and

328

4 at locations where κm·δL. −0.3; the preheat zone can

329

be thickened by nearly a factor of four. For highly

nega-330

tively curved shapes (i.e. Phase 3), the methane preheat

331

zone is more broadened compared to the propane flame.

332

The preheat layer thickening in highly curved regions

333

occurs near the burner centerline, whereas the weakly

334

curved edges of the flame remain relatively unaffected

335

by the FVI.

336

Because the most significant changes are observed in

337

the highly curved region at the centerline, the reaction

338

zone thickness is examined at this location as a function

339

of time. The reaction zone thickness δrec is defined as

340

the distance between c= 0.7 and c = 0.85 at r = 0 and

341

shown versus the local curvature in Fig. 5(b). The time

342

evolution manifests as first decreasing and then

increas-343

ing curvature values. For large negative curvatures, the

344

reaction zone is strongly thickened for the Lee f f > 1

345

propane/air flame, but only slightly thickened for the

346

Lee f f ≈ 1 methane/air flame.

347

Note that Phase 2 and the early portion of the

inter-348

action correspond to times of high strain rate. However,

349

the preheat and reaction zone thicknesses are only

min-350

imally altered during this time. Hence, the flame

thick-351

ness is much more sensitive to curvature than strain.

352

Moreover, for both the preheat and reaction zones,

353

structural changes are noticeable only when the radius

354

of curvature is smaller than twice the laminar flame

355

thickness.

356

Previous conclusions pertain to the thermal structure

357

of the flame. We now turn our attention to some

in-358

dicators of the kinetic activity of the reaction zone by

359 0.2 0.6 1 1.4 1.8 -1.8 -1.4 -1 -0.6 -0.2 0.2 0 1 2 3 Increasing time

Figure 5: (a) Preheat zone thickness vs normalized curvature along the flame contour for four different phases (each phase is represented by a specific symbol). (b) Normalized reaction zone thickness evaluated at the flame tip during the whole FVI. Curvature is normalized with the laminar flame thermal thickness. Red (filled) and black (open) colors (symbols) represent C3H8/air and CH4/air mixtures. The arrow

indicates the direction of increasing time.

analysing the chemiluminescence signals of OH* and

360

CH*.

361

3.3. CH* and OH* measurements

362

Given the above-noted changes to the reaction zone

363

thickness in highly negatively curved flame segments, it

364

is useful to further investigate the thermal and chemical

365

processes occurring in this region. CH* and OH* are

366

two reasonable metrics representing the reaction zone.

367

However, meaningful interpretation of

chemilumines-368

cence measurements requires knowledge of the

kinet-369

ics leading to the formation of the excited state. The

370

reaction of ethynyl radical C2H with atomic oxygen is

371

the dominant source of CH* formation (C2H + O ↔

372

CH* + CO) [28], whereas the emission intensity of

373

OH* is mainly proportional to the local concentration

374

of CH and O2(CH+ O2 ↔ OH*+ CO) for

hydrocar-375

bon flames [29]. The ratio OH*/CH* gives an indication

376

of the balance of these two mechanisms.

377

Figure 6 displays Abel deconvoluted images of CH*,

378

OH* and OH*/CH* for a curved propane flame at Phase

379

3. Elevated CH* concentrations occurred at the flame

380

tip, whereas the OH* concentration was relatively

con-381

stant along the flame; the OH*/CH* ratio decreased near

382

the flame tip. The relative decrease of OH* compared to

383

CH* could be due to several factors. In Refs. [30, 31],

384

it was found that OH*/CH* ratio was inversely

propor-385

tional to equivalence ratio for lean and stoichiometric

386

(8)

0 200 400 600 0 2 4 104

0 100 200

Figure 6: Snapshots of CH*, OH* and OH*/CH* (left to right) for a curved propane flame at Phase 3.

flames. However, differential diffusion of propane and

387

oxygen would be expected to result in leaner conditions

388

at the flame tip, which is counter to the experimental

389

results.

390

Recall that the lean propane flame is characterized by

391

Lee f f > 1, meaning that thermal diffusion is dominant

392

compared to molecular diffusion. This suggests that the

393

increase in CH* intensity in regions of high negative

394

curvature may be attributed to higher thermal di

ffusiv-395

ity enhancing the formation of CH* precursors. This

396

was verified numerically using a counterflame

config-397

uration, which showed that preheating fresh reactants

398

increases the concentration of O, C2H, and CH*. In

399

contrast, the concentration of OH* was relatively

un-400

changed because the majority of O2is consumed before

401

reaching the OH* layer. Hence, the increase in CH*

402

and decrease in OH*/CH* near the flame tip is due to

403

preferential thermal diffusion in this region.

404

To further quantify the effect of local curvature on

405

chemiluminescence, CH* and OH* intensities were

in-406

tegrated normally to the c = 0.7 iso-contour and then

407

normalized by a reference value corresponding to the

408

flat flame. These are plotted versus the local flame

cur-409

vature in Fig. 7. For the propane flame, both CH* and

410

OH* were negatively correlated with curvature, though

411

with much greater sensitivity for CH* in the range

412

−0.6. κm·δL. −0.2. Beyond κm·δL. −0.6, neither

413

OH* nor CH* were sensitive to curvature.

414

A similar behavior has been identified for the effect of

415

curvature on the local burning velocity [17, 18, 32];

neg-416

atively curved flame segments have higher local flame

417

speeds for Lee f f > 1, whereas, the correlation is

op-418

posite for Lee f f < 1. The curve representing the local

419

burning velocity as a function of normalized curvature

420

for a φ = 0.7 propane flame, provided by Bell et al.

421

[32], is also shown in Fig. 7. As can be seen from

422

this figure, the slopes of the three curves are different.

423

This means that CH* and OH* radicals are not directly

424

correlated to local burning velocity or heat release rate

425

when the flame is curved. There might be other

pa-426

rameters to consider in order to deduce the right

de-427

pendence between chemiluminescent species and heat

428

release rate. Hence, CH* and OH* species can be

inter-429

preted as qualitative indicators of heat release rate and

430

cannot be used to quantify it.

431

Concerning the methane flame, it was found that CH*

432

and OH* intensities are not strongly affected by

curva-433

ture and remain nearly constant (see Fig. 7). One would

434

expect that mixtures with Lee f f ≈ 1 will have such a

be-435

havior. It is important to note that intermediate species

436

as atomic hydrogen and atomic oxygen may also have a

437

significant impact, particularly for flames characterized

438 by Le < 1. 439 -1.8 -1.4 -1 -0.6 -0.2 0.2 0.8 1 1.2 1.4 1.6 1.8 2 0.8 1 1.2 1.4 1.6 1.8 2

Figure 7: Normalized and integrated values of the CH* (red symbols) and OH* (black symbols) intensities along the flame contour vs the normalized curvature. Filled and open symbols stand for propane and methane flames respectively. Overlaid dashed curve is reproduced from the results presented in Bell et al. [32].

4. Concluding remarks

440

The influence of stretch and curvature on the

struc-441

ture of different Lewis number flames have been

ex-442

plored in the particular configuration of a flame/vortex

443

interaction. The imbalance between thermal and

reac-444

tant diffusion is evaluated by comparing a

stoichiomet-445

ric methane flame to a lean propane flame. Analysis of

446

the thermal structure and the kinetic activity is

under-447

taken on the basis of Rayleigh scattering together with

448

CH* and OH* chemiluminescence measurements,

re-449

spectively.

450

In the present case, the vortex intensity and size

rel-451

ative to the flame properties lies within the wrinkled

452

flame regime of Refs. [1, 2]. Therefore, one should have

453

expected only marginal structural variations of the flame

454

zone. Moreover, it is often postulated that even though

455

the preheat zone might be strongly influenced by

hydro-456

dynamic straining and curvature, the reaction zone is

457

much more robust and experiences only marginal

thick-458

ness variations. Our study demonstrates that these two

(9)

statements are violated even at moderate vortex

inten-460

sity and large vortex size. Indeed, a significant thinning

461

(thickening) is observed during the first (second) period

462

of the interaction where the strain rate (curvature)

con-463

tribution to stretch dominates. This reinforces the need

464

for distinguishing strain and curvature effects as in Refs.

465

[10–12] for the flame speeds. More importantly, by

dif-466

ferentiating the flame zone into a preheat and reaction

467

zone, it is clearly emphasized that the structure of the

468

reaction zone can be also profoundly altered. The

thick-469

ening of the reaction was observed only when the flame

470

radius of curvature was of the order of the flame

thick-471

ness and for a non unity Lewis number flame. These

ob-472

servations are further confirmed by analysing CH* and

473

OH*, i.e. two reasonable markers of the reaction zone

474

activity. We show that the reaction zone of a propane

475

(methane) flame is significantly (marginally) sensitive

476

to curvature.

477

As an overall conclusion, temperature fields together

478

with the chemiluminescence signals of CH* and OH*

479

highlight that in contrast with common statements, a

480

moderate vortex intensity with typical size much larger

481

than the flame thickness is sufficient to induce

signif-482

icant changes in the flame structure. Although FVIs

483

should be interpreted as an intermediate situation

be-484

tween stretched laminar and turbulent flames, the

cur-485

rent results suggest that the inner flame flame structure

486

may be modified at lower Karlovitz numbers than

clas-487

sically expected.

488

Acknowledgments

489

The authors gratefully acknowledge the financial

490

support provided by the French Government

Pro-491

gram "Investissements d’avenir" through the LABEX

492

CAPRYSSES ANR-11-LABX-0006-01, NSERC

493

Canada through Grant RGPIN 2017-06501, and the

494

Ontario Ministry of Research and Innovation Early

495

Researcher Award program. We are also thankful to the

496

CNRS and the University of Orléans.

497

References

498

[1] T. Poinsot, D. Veynante, S. Candel, J. Fluid Mech. 228 (1991)

499

561–606.

500

[2] W. L. Roberts, J. F. Driscoll, M. C. Drake, L. P. Goss, Combust.

501

Flame 94 (1993) 58–69.

502

[3] O. Colin, F. Ducros, D. Veynante, T. Poinsot, Phys. Fluids 12

503

(2000) 1843–1863.

504

[4] F. Charlette, C. Meneveau, D. Veynante, Combust. Flame 131

505

(2002) 159–180.

506

[5] S. Bougrine, S. Richard, O. Colin, D. Veynante, Flow, Turbul.

507

Combust. 93 (2014) 259–281.

508

[6] F. Thiesset, G. Maurice, F. Halter, N. Mazellier, C. Chauveau,

509

I. Gökalp, Proc. Combust. Inst. 36 (2017) 1843–1851.

510

[7] J.-M. Samaniego, T. Mantel, Combust. Flame 118 (1999) 537–

511

556.

512

[8] T. Mantel, J.-M. Samaniego, Combust. Flame 118 (1999) 557–

513

582.

514

[9] C. Mueller, J. Driscoll, D. Reuss, M. Drake, Proc. Combust.

515

Inst. 26 (1996) 347–355.

516

[10] F. Thiesset, F. Halter, C. Bariki, C. Lapeyre, C. Chauveau,

517

I. Gökalp, L. Selle, T. Poinsot, J. Fluid Mech. 831 (2017) 618–

518

654.

519

[11] J. Bechtold, M. Matalon, Combust. Flame 127 (2001) 1906–

520

1913.

521

[12] G. K. Giannakopoulos, A. Gatzoulis, C. E. Frouzakis, M.

Mat-522

alon, A. G. Tomboulides, Combust. Flame 162 (2015) 1249–

523

1264.

524

[13] C. Law, S. Ishizuka, P. Cho, Combust. Sci. Technol. 28 (1982)

525

89–96.

526

[14] M. Mizomoto, Y. Asaka, S. Ikai, C. Law, Proc. Combust. Inst.

527

20 (1985) 1933–1939.

528

[15] T. Poinsot, T. Echekki, M. Mungal, Combust. Sci. Technol. 81

529

(1992) 45–73.

530

[16] T. M. Vu, M. S. Cha, B. J. Lee, S. H. Chung, Combust. Flame

531

162 (2015) 1614 – 1621.

532

[17] D. Haworth, T. Poinsot, J. Fluid Mech. 244 (1992) 405–436.

533

[18] C. Rutland, A. Trouvé, Combust. Flame 94 (1993) 41–57.

534

[19] C. M. Vagelopoulos, J. H. Frank, Proc. Combust. Inst. 30 (2005)

535

241–249.

536

[20] V. Dribinski, A. Ossadtchi, V. A. Mandelshtam, H. Reisler, Rev.

537

Sci. Instrum. 73 (2002) 2634–2642.

538

[21] G. P. Smith, D. M. Golden, M. Frenklach, N. W. Moriarty,

539

B. Eiteneer, M. Goldenberg, C. T. Bowman, R. K. Hanson,

540

S. Song, W. C. Gardiner Jr, et al., Gas Research Inst. (1999).

541

[22] C.-W. Zhou, Y. Li, E. O’Connor, K. P. Somers, S. Thion,

542

C. Keesee, O. Mathieu, E. L. Petersen, T. A. DeVerter, M. A.

543

Oehlschlaeger, et al., Combust. Flame 167 (2016) 353–379.

544

[23] W. Gardiner, Y. Hidaka, T. Tanzawa, Combust. Flame 40 (1981)

545

213–219.

546

[24] J. Fielding, J. H. Frank, S. A. Kaiser, M. D. Smooke, M. B.

547

Long, Proc. Combust. Inst. 29 (2002) 2703–2709.

548

[25] J. A. Sutton, J. F. Driscoll, Opt. Lett. 29 (2004) 2620–2622.

549

[26] C. Law, C. Sung, G. Yu, R. Axelbaum, Combust. Flame 98

550

(1994) 139–154.

551

[27] J. F. Driscoll, Prog. Energy Combust. Sci. 34 (2008) 91–134.

552

[28] K. Devriendt, J. Peeters, J. Phys. Chem. A 101 (1997) 2546–

553

2551.

554

[29] S. Carl, M. Van Poppel, J. Peeters, J. Phys. Chem. A 107 (2003)

555

11001–11007.

556

[30] J. Kojima, Y. Ikeda, T. Nakajima, Combust. Flame 140 (2005)

557

34–45.

558

[31] M. Orain, Y. Hardalupas, CR Mecanique 338 (2010) 241–254.

559

[32] J. B. Bell, R. K. Cheng, M. S. Day, I. G. Shepherd, Proc.

Com-560

bust. Inst. 31 (2007) 1309–1317.

561

Figure

Figure 1: Optical configuration. CL: cylindrical lens, DM x : dichroic mirror, SL x : spherical lens, F x : bandpass filter.
Figure 2: Time sequence of temperature field for propane flame at φ
Figure 4: Normalized |∇c| profiles versus progress variable c at the flame centerline for different phases during FVI for P09 case (filled symbols) and M10 case (open symbols).
Figure 7: Normalized and integrated values of the CH* (red symbols) and OH* (black symbols) intensities along the flame contour vs the normalized curvature

Références

Documents relatifs

tinine serait plus sensible que celle du DFG pour détec- ter la NPC, mais la corrélation clinique de cette variation d’incidence reste à évaluer. En effet, près de la moitié

We now consider the response of the climatic oscillator described by eqs. Our purpose is to show that the forcing provides the synchroni- zing element that was

Peu d ’études citent la prise en charge et le devenir chez des nourrissons hospitalisés en unité de réanimation pédia- trique [3], avec des spécificités qui leur sont propres :

II Stabilization of confined premixed CH 4 /H 2 /air swirling flames 139 4 Analysis of topology transitions of swirled flames interacting with the combustor peripheral wall 141

Infection with the protozoan Toxoplasma gondii is one of the most frequent parasitic infections worldwide. This obligate intracellular parasite was first described

The mechanism has been written to reproduce the formation and consumption of formic, acetic, propionic and propenic acids under flame conditions; i.e.. the low- temperature

Keywords: parameters estimation, maximal conductance ionic parameters, bidomain model, optimal control with PDE constraints, first order optimality conditions, physiological

Dans cette étude rétrospective, nous avons analysé des éléments épidémiologiques, des données cliniques et thérapeutiques surtout symptomatiques de quelques