particle streak velocimetry

8

Click here to load reader

Upload: hyperliz

Post on 29-May-2017

215 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Particle Streak Velocimetry

Particle streak velocimetry and CH laser-inducedfluorescence diagnostics in strained, premixed,

methane-air flames

J. M. Bergthorson†∗, D. G. Goodwin‡, and P. E. Dimotakis†

† Graduate Aeronautical Laboratories,‡ Department of Mechanical Engineering,

California Institute of Technology, Pasadena, CA, USA, 91125

Abstract

We present the use of simultaneous Particle Streak Velocimetry (PSV) and CH Planar Laser Induced Fluorescence(PLIF) diagnostics in the study of planar, strained, premixed, methane-air flames, stabilized in a jet-wall stagnationflow. Both PSV and PLIF data are imaged at high spatial resolution and sufficiently high framing rates to permit anassessment of flame planarity and stability. Concurrent measurements of mixture composition, (Bernoulli) static-pressure drop, and stagnation-plate temperature provide accurate boundary conditions for numerical simulations.The new PSV implementation is characterized by very low particle loading, high accuracy, and permits shortrecording times. This PSV implementation and analysis methodology is validated through comparisons withprevious laminar flame speed data and detailed numerical simulations. The reported diagnostic suite facilitatesthe investigation of strained methane-air flames, as a function of nozzle-plate separation to jet-diameter ratio,L/d, and equivalence ratio, Φ. Flames are simulated using a one-dimensional streamfunction approximation,with full chemistry (GRI-Mech 3.0), and multi-component transport. In general, we find good agreement betweenexperiments and simulations, if boundary conditions are specified from measured velocity fields. Methane-airflame strength appears to be slightly overpredicted, with the largest disagreements for lean flames.

Keywords: stagnation, one-dimensional, flame speed, PLIF, Cantera

1. Introduction

Planar, strained flames provide a valuable devel-opment and validation testbed for transport and ki-netics models that can then be used in any combus-tion environment. The high parametric dimensional-ity of these models requires many detailed and ac-curate experiments over a sufficiently large range ofconditions. Such experiments are enhanced by diag-nostics that simultaneously permit detailed flow andchemical-species data to be measured and comparedto model predictions. This paper describes a suite ofdiagnostics developed for this purpose and employedto probe planar, strained, methane-air flames, stabi-

∗Corresponding author.E-mail address: [email protected]

lized in a jet-wall stagnation flow. This setup yieldsa flow with boundary conditions that can be accu-rately specified, facilitating simulation and compar-isons with experiment. This flow can also, with care,be stable to high Reynolds numbers. The diagnos-tics are optimized for accuracy, minimal flame distur-bance, and rapid simultaneous recording of flow ve-locity and CH radical profiles.

Diagnostics documented here include extensionsto and improvements of Particle Streak Velocime-try (PSV), complemented by simultaneous CH Pla-nar Laser Induced Fluorescence (PLIF) imaging at10 Hz. PLIF spectroscopyconfirms the CH excitationand helps optimize dye-laser wavelength for signal-to-noise ratio. PLIF images are processed to extractCH radical profiles and assess flame planarity. The10 fps PLIF sequence permits an assessment of flame

Page 2: Particle Streak Velocimetry

stability under all conditions. Only data from stableflame conditions are included.

Particle Streak Velocimetry (PSV) [1–3], a tech-nique similar to particle tracking velocimetry [4], isused for instantaneous flow-field measurements and,in particular, axial velocities along the flow center-line. In combustion, PSV was previously used forqualitative flow descriptions (e.g., Sugiyama [5]). Ex-tensions and improvements to PSV implemented hereinclude digital imaging and processing, and new anal-ysis techniques. The implementation yields a La-grangian measurement of velocity that, in principle,requires only a single particle traversing a steadyflame. The consequentPSV accuracy competes favor-ably with LDV and PIV techniques, while a 1-2 orderof magnitude reduction in particle loading minimizesflame disturbances, and Mie-scattering and stray-lightinterference in PLIF images. Velocity data are used tospecify boundary conditions for simulations and arecompared to predictions.

The spatial distribution of the CH radical can bemeasured with PLIF. Its narrow spatial profile is well-correlated with flame location and provides a sensitivetest of direct predictions of strained-flame models.

Simultaneous measurements of air, fuel, and dilu-ent mass fluxes, as well as of stagnation plate tem-perature, allow an accurate specification of boundaryconditions for simulations.

Experimental profiles are compared to one-dimensional simulation predictions, using the Can-tera software package [6]. The simulations in-corporate GRI-Mech 3.0 kinetics [7] with multi-component transport [8]. The simulations rely ona one-dimensional streamfunction formulation [9].Few comparisons between this formulation and ac-tual velocity data have been documented, however.Law and coworkers studied methane-air, opposed-jetflames for lean, stoichiometric, and rich mixtures, us-ing LDV and CARS for velocity, temperature, andmajor-species measurements to quantify the effect ofstretch on flame structure [10–13]. To compare exper-imental and simulated data, a potential-flow bound-ary condition with a variable inflow mass flux is usedto visually match the profiles. The authors reportgeneral agreement for temperature and major speciesprofiles when flame location is adjusted to match themeasurements.

The new PSV methodology is validated by com-paring extrapolated laminar flame speeds with exist-ing data and numerical simulations. Slightly lean,strained, methane-air flames are studied as a functionof the nozzle-stagnation plate separation distance, L,to assess the simplified hydrodynamic model. Flametemperature dependence is studied by mixture dilu-tion with excess nitrogen. The diagnostics are appliedto methane-air flames, under similar strain-rate condi-tions, as a function of equivalence ratio, Φ. The ap-proach and diagnostics permit an assessment of thenumerical simulation predictions of strained-flames.

2. Experiments

A room-temperature, atmospheric-pressure jet isgenerated from a contoured nozzle with an exit di-ameter of d = 9.9 mm that impinges on a constant-temperature (water-cooled) copper stagnation plate.Three K-type thermocouples are embedded on thecenterline, spaced vertically between the stagnationand cooled surface, to allow monitoring of wall tem-perature and temperature gradients.

The pressure difference between the jet plenuminterior and a point just outside the jet-coreflow region is measured with a 1 torr full-scaledifferential-pressure transducer (BOC EdwardsW57401100 and W57011419). The Bernoulli

velocity, UB =√

(2∆p)/(ρ[1 − (d/dP)4]), is thencalculated, where ∆p is the nozzle static pressuredrop, ρ the fluid density, d the nozzle exit diameter,and dP the plenum (inner) diameter. The Bernoullivelocity will be higher than the jet-exit velocity forL/d ≤ 1, owing to a velocity defect produced bystreamline curvature at the nozzle-exit [14].

Fuel, oxidizer, and nitrogen-diluent mass flowrates are set using sonic metering valves and mon-itored concurrently (Omega FMA868-V-Methane,FMA872-V-Air and FMA873-V-Air, calibrated usinga Bios DryCal ML-500). Estimated uncertainty in themass-flow measurement of the air and fuel streams is,approximately, 1%, each. Bernoulli pressure, mass-flow, and temperature data acquisition is synchro-nized with digital-image acquisition, allowing accu-rate specification of simulation boundary conditions.

2.1. Particle Streak Velocimetry (PSV)

Flow velocities along the jet centerline are mea-sured using Particle Streak Velocimetry (PSV) [1–3]. The implemented PSV methodology yields low-fractional-error axial-velocity data. The resultingaccuracy is comparable to that obtained with LDVor PIV, while offering several advantages in flameenvironments. The reduced particle loading re-quired for PSV minimizes flame disturbances intro-duced by particle seeding that can alter the effectivethermal/heat-capacity environment, or the chemicalkinetic/catalytic environment by providing surface-chemistry sites. Low PSV particle loading also re-duces Mie-scattering interference in CH PLIF image-data, improving CH profile statistics. Particle load-ing required for accurate velocity measurements withPSV is an order of magnitude lower, or more, than re-quired for LDV, or PIV. In a single PSV image frame,one or two particles traversing the vertical extent ofthe imaged region can suffice. In contrast, PIV mea-surements require a dispersion of particles throughoutthe domain in any one (short-time) exposure. WithLDV, high particle number densities are required forconverged statistics in a reasonable time.

A single PSV image captures the entire veloc-ity field, making it ideal for short-run-time exper-iments. An example PSV image for a stagnationflame is shown in Fig. 1a. In axisymmetric, steady

Page 3: Particle Streak Velocimetry

stagnation flow, axial velocities along the centerlinecan be reliably measured; particle paths do not crossor overlap, and occasional out-of-plane particle dis-placements are small and easily discernible (in-/out-of-focus streaks). The high sensitivity of the scat-tering cross section to particle size, in the size rangeemployed, allows easy identification of agglomeratesthat may not track high spatial-gradient regions inthe flow. Streaks used for PSV processing are fromin-plane, non-agglomerated particles. In a variable-velocity field, particles track the flow if the dimen-sionless product of the local strain rate, σ = ∂u/∂x,and the Stokes time, τS ≡ ρpd

2p/(18µ), is small,

i.e., provided στS � 1. The measurements reliedon alumina particles (Al2O3; median size, dp

∼=0.8µm, ρp

∼= 3830 kg/m3; Baikowski Malakoff,RC-SPT DBM). For the strain rates in these experi-ments, σ τS ≤ 3 × 10−3.

A Coherent I-90 Ar+ laser, operated at 2 − 3 W,provides the PSV illumination source. Two cylin-drical lenses generate a thin laser sheet (≈ 200µm)in the field of view. An Oriel (Model 75155)chopper with a 50% duty-cycle wheel modulatesthe laser beam. The chopper wheel is placed ata laser-beam waist to minimize on-off/off-on transi-tion times. Chopping frequencies are in the range,1.6kHz ≤ νc ≤ 2.4 kHz, with νc optimized depend-ing on flow velocity.

PSV image data are recorded using the in-house-developed “Cassini” digital-imaging system that re-lies on a low-noise, 10242-pixel CCD, with a (square)12µm pitch. A PSV image frame rate of 4 fps ischosen to further minimize particle-loading require-ments. Magnification ratios are in the range of 1:1.0 –1:1.5 using a Nikon 105 mm, f/2.8 macro lens (witha 514.5 nm bandpass filter).

Local velocities, u(x), are estimated from streakpairs as, u(x) ∼= ∆X(x)/∆t, yielding uI = LI/τcand uII = LII/τc, where τc = 1/νc (reciprocalof chopper frequency) and L I = x2s − x1s andLII = x2e − x1e are the distances from the start/endof one streak to the start/end of the next, respectively(cf. Fig. 1b). Using the same intensity threshold ona streak pair removes systematic errors in applyingthe Lagrangian time interval, τc. This methodologyis spatially second-order accurate and produces goodagreement between velocity values derived from eachstreak pair. Streak lengths are estimated using bi-cubic fits on the 2-D streak-intensity image data, sam-pled to a 0.1-pixel resolution in both dimensions. Anintensity threshold of ≈ 0.3 of the maximum inten-sity of each streak is used. The results do not dependon this choice and yield an rms error of ≈ 0.01U j,where Uj is the jet-exit velocity.

The velocity estimate, uI, is located at xI = (x1s+x2s)/2 + (w1 +w2)/4, where xis is the spatial loca-tion of the start of the ith streak and wi is the widthof the ith streak (cf. Fig. 1b). Similarly, uII is locatedat xII = (x1e + x2e)/2 − (w1 + w2)/4, where xie

is the location of the end of the ith streak. PSV spa-

tial resolution is comparable to that of other particlevelocimetry techniques, e.g. PIV, LDV, that rely oncomparable spatial displacements (e.g., 10 − 30 pix,or ≈ 100 − 300µm, for this flow). These methodsmeasure the distance traveled by a particle along aLagrangian path in a fixed time (PIV, PSV), or thetime required to traverse a number of fringes in space(LDV). Particle methods require care in regions ofhigh fractional change in speed along individual parti-cle track segments, here limited to measurements veryclose to the wall, or in regions of high velocity curva-ture.

2.2. Planar Laser Induced Fluorescence (PLIF)

CH fluorescencemeasurements are obtained by ex-citing the Q1(7) transition of the B2Σ −X2Π(0, 0)CH band at 390.30 nm, in the UV [15]. The UV beamis obtained from a tripled Nd:YAG (355 nm)-pumpeddye laser (Spectra-Physics PRO-290 & Sirah Preci-sionScan). Pulse duration is ≈ 5 ns with a linewidthof ≈ 0.1 cm−1 (1.5 pm). Fluorescence is recordedfrom theA−X(1, 1),A−X(0, 0), andB−X(0, 1)bands in the 420 − 440 nm range. Excitation to theB state yields a higher signal than excitation to the Astate [16] and a large wavelength shift betweenexcita-tion and fluorescence, facilitating filtering of Mie- andRayleigh-scattering interference, which is importantin particle-seeded flames. TheQ1(7) band provides ahigh signal level and low temperature sensitivity.

The output of the dye laser is passed through a pairof cylindrical lenses (150 mm and 500 mm at right an-gles) and yields a laser sheet with a Rayleigh rangecentered on the jet axis. Fluorescence is collectedwith a Nikkor 50 mm, f/1.2 lens at magnificationsnear 1:1, with a Schott KV-418 long-pass filter toreject scattering of the UV laser illumination, whiletransmitting approximately 90% of the fluorescencenear 430 nm. Detection relies on a lens-coupled inten-sifier (Ultra-Blue Gen-III, Cooke Corp. VS-364) witha cooled CCD (Cooke Corp. SensiCAMQE, binnedto 344 × 260 pix2; a binned pixel maps into 46µmin the flow). A 70 ns gate time rejects chemilumines-cence while retaining fluorescence. For the PLIF pro-files presented here, a laser power of ≈ 10 mJ/pulseresults in a saturated laser spectral intensity of Iν =Ep/(τpAbSW ) ≈ 108 (W/cm2)/cm−1, where Ep

is the pulse energy, τp the pulse length, Ab the laserbeam cross-sectional area, and SW the laser spectralwidth [17]. A sample CH PLIF (single-exposure) im-age is reproduced in Fig. 2a, and an average of 1000exposures is shown in Fig. 2b. To correct for darkcurrent and Rayleigh scattering, an averaged image ofthe laser sheet (no flame) is subtracted from the imagedata. CH profiles are obtained from single-shot im-ages by summing across the (flat) central portion ofthe flame. Single-image 60-column averaging, aboutthe jet axis, yields good profile statistics.

Figure 3 shows an experimentally measured CHexcitation scan, at a resolution of 0.5 pm. Spectrumsimulations are performed with LIFBASE [18]. This

Page 4: Particle Streak Velocimetry

scan was recorded with a laser power of 0.2 mJ/pulse,in a 1mm×2cm sheet, producing a partially saturatedlaser spectral intensity of Iν ≈ 2 × 106 ≈ Isat

ν ≈106 (W/cm2)/cm−1.

3. Numerical method

Stagnation flame simulations are performed us-ing the Cantera reacting-flow software package [6].The one-dimensional solution for stagnation flows ap-proximates the flow in terms of a streamfunction,ψ(x, r) = r2U(x), with U(x) = ρu/2, where u isthe axial velocity [9]. The axisymmetric momentumequation then becomes,

2 Ud

dx

(1

ρ

dU

dx

)−1

ρ

(dU

dx

)2− d

dx

d

dx

(1

ρ

dU

dx

)]= Λ . (1)

In this formulation, Λ = (1/r) dp/dr must be aconstant. By treating Λ as unspecified, four bound-ary conditions are imposed on this third-order ordi-nary differential equation at x = 0 and x = �,with 0 < � ≤ L a suitably chosen interior point,permitting U(0) = 0, U ′(0) = 0, U(�) = U� ,U ′(�) = U ′

�. Exploiting the (cold) outer-flow so-lution to Eq. 1, which is a parabola, a quadratic isfit to the cold flow data upstream of the flame. U �

and U ′� are then calculated from the fit, at x = �,

minimizing errors that could be introduced by an in-consistent specification of flow boundary conditions,or from differentiation of the raw data. Energy andspecies equations are also solved with specification ofinlet composition, inlet temperature, and stagnation-wall temperature. A no-flux boundary condition forspecies is applied at the wall. The simulations use amulti-component transport model and the GRI-Mech3.0 kinetics mechanism [7, 8]. Simulations of laminarflame speeds are performed using a freely propagat-ing flame code with multi-component transport andthe same kinetics mechanism [19].

4. Results and discussion

The PSV methodology is validated by comparingextrapolated laminar flame speed measurements withprevious data and numerical simulations. Figure 4aplots sample reference flame speed and strain-ratedata, and the extrapolated flame speed. Vertical errorbars are based on linear regression techniques. Hor-izontal error bars represent the estimated uncertaintyin Φ of 1.4%. Data at equivalence ratios of Φ = 0.75,0.84, and 0.95 are given in Fig. 4b, and compared tonumerical simulations and previous LDV-based data[20, 21]. The measurements agree with previous dataand, collectively, indicate lower flame speeds thanpredicted.

The diagnostic suite is applied to methane-airflames as a function of scaled separation distance,L/d, and Φ. Table 1 lists the equivalence ratio, scaledseparation distance, and dilution for the experimentsreported here. The first subset of experiments is atΦ ∼= 1.0, with variable L/d (Table 1a). A diluted

methane-air flame is also studied (Table 1b). Flamechemistry effects are probed at L/d = 0.8 by varyingequivalence ratio and dilution (Table 1c). The dilutionlevel is chosen to maintain similar stretch conditionsas mixture-fraction is varied.

Figure 5 shows the results for a near-stoichiometric(Φ = 0.98) methane-air flame, at L/d = 0.6. CHprofiles, obtained from 60-column averages in single-exposure images, are normalized by peak values. Thesimulation predicts flame location with good agree-ment in both the flowfield and CH profile. Ther-mophoresis will cause a particle drift in the hightemperature-gradient region of the flow [11]. Utiliz-ing the simulated temperature profile, thermophoreticcorrections yielded a maximum error of 3% in thehigh gradient region of the velocity profile (preheatzone). Velocity data in this region are used to visuallyassess simulation predictions and these (small) ther-mophoretic corrections are neglected; velocity data inthe cold-flow region, where a quadratic is fit to de-termine flow boundary conditions, are unaffected bythermophoresis.

To compare simulations with experiments, the dif-ference between measured and predicted peak CHconcentration location, xCH,PLIF andxCH,sim, scaledby the simulated CH layer thickness, δCH, is calcu-lated. CH layer thicknesses are calculated as the FullWidth at Half Maximum (FWHM) of Gaussian fits tosimulated profiles. The difference between xCH,PLIF

and xCH,sim is less than 0.1 δCH, in this case.The scaled separation distance, L/d, is an impor-

tant flow parameter. Figure 6 shows the results for anear-stoichiometric (Φ = 0.98) methane-air flame, atL/d = 0.8. Predicted and experimental peak CH lo-cations agree within 0.2 δCH for L/d = 0.6, 1.0 and1.2, and within 2 δCH for L/d = 0.8, at Φ ∼= 1.0. Ex-perimental and predicted CH peak locations exhibitreasonable agreement at Φ ∼= 1.0, indicating goodprediction of strained flame location for a variety ofimposed hydrodynamic fields. The variation of flowconditions at constant Φ indicates that the simplifiedhydrodynamic model can capture the experiment. Forthe remainder of this study, the scaled separation dis-tance is fixed at L/d = 0.8.

To investigate flame-temperature effects, a dilutedΦ = 0.98 flame is studied. Excess nitrogen is addedto the premixture to reduce the flame temperature.Reasonable agreement was found between measuredand predicted velocity and CH profiles, indicating thatthe simulations can capture variations in flame heatrelease and attendant temperature-dependent trans-port effects. The difference between simulated andmeasured CH peak locations is less than 2 δCH.

Flame chemistry effects are probed by varying Φ.Representative results are given in Figs. 7 and 8 thatcompare experimental and predicted profiles for alean and rich flame, respectively. For the lean flame(Φ = 0.69), the predicted CH-profile is located up-stream of the experimental one, consistent with anoverprediction of strained-flame speed. Correspond-ingly, post-flame velocity is higher than measured.

Page 5: Particle Streak Velocimetry

For the rich flame (Φ = 1.31), good agreement isseen between experiment and simulation. Far fromstoichiometry, the results are sensitive to inlet compo-sition. A simulated velocity profile with Φ decreasedby 1.4% (the direction required for agreement) is alsoincluded in Fig. 7. The 1.4% increment represents theestimated maximum uncertainty in Φ. For lean andrich flames, simulations exhibit high sensitivity to theinlet composition and its measurement uncertainty.

Results for methane-air flames over a range ofequivalence ratios are presented in Fig. 9. In anattempt to reduce the number of parameters variedbetween experiments, strong burning flames (Φ ∼=1.1, 1.2) are diluted with excess nitrogen to maintainan approximately constant flame temperature [22].This permits a similar strain-field to be established asequivalence ratio is varied, allowing some decouplingof the various effects. We note reasonable agreementfor methane-air flames, with a maximum scaled errorbetween measured and predicted CH peak locationsof ≈ 2.3 δCH for the leanest flame (cf. Fig. 9a). Errorbars represent the results from simulations with Φ in-creased and decreased by its maximum estimated un-certainty of 1.4%. For the richest flame (Φ = 1.31), asolution could not be found with Φ increased by 1.4%and the error bar represents a decrease of 1.4% in Φ.

Figure 9b compares experimental and simulatedCH profile thicknesses, referenced to the Φ = 0.96flame. Measured CH-profile thickness is an ensembleaverage of the FWHM of individual Gaussian fits tosingle-exposure profiles (60-column average over theflat, central portion of the flame). For the Φ = 0.96,L/d = 0.8 flame, the measured CH-profile thicknessis approximately twice the simulated value. The PointSpread Function (PSF) of the imaging system and thetrue CH profile can be approximated by Gaussians.A measured PLIF profile will then also be a Gaus-sian that is a convolution of the two, with a compositewidth (squared), δ2PLIF, that is the sum of the PSF,δ2PSF, and true CH profile, δ2CH, widths squared, i.e.,δ2PLIF

∼= δ2PSF+δ2CH. PSF width squared is estimatedbased on the undiluted, Φ = 0.96, L/d = 0.8 flameby subtracting the measured and simulated widthssquared. This PSF width is systematically applied tostudy the difference between predicted and measuredreference profile thicknesses as a function of equiva-lence ratio. Figure 9b plots the mean FWHM calcu-lated from multiple shots, 700 < n ≤ 1000, and errorbars (one standard deviation). Simulations underpre-dict variations in δCH with Φ and dilution. Beam-steering, lensing, or profile-broadening artifacts areestimated not to be significant, or Φ-dependent, andcannot account for this discrepancy.

5. Conclusions

Combined PSV and CH PLIF diagnostics are pre-sented that yield high-accuracy measurements of ve-locity fields and of the spatial extent of relative CH-concentration profiles. The new PSV implementationrequires very low particle loading, resulting in mini-

mal flame disturbances and improving statistics in si-multaneous PLIF measurements. This implementa-tion and analysis methodology is validated throughgood agreement with previous laminar flame-speeddata and detailed numerical simulations. Numericalsimulations overpredict laminar flame speeds for leanflames.

The diagnostic suite facilitates investigation ofstrained premixed flames, performed here formethane-air mixtures, as a function of the nozzle-plate separation and equivalence ratio. Axial ve-locity and CH profiles are extracted from PSV andPLIF field data and compared to one-dimensionalsimulations. Strained flames are simulated us-ing a one-dimensional streamfunction approximation,with detailed chemistry (GRI-Mech 3.0) and multi-component transport. The diagnostics permit a com-plete boundary-condition specification to the one-dimensional flow-transport-kinetic model, allowingdetailed comparisons of measured and predicted ve-locity and CH profiles in strained stagnation flames.

Near-stoichiometric, strained, methane-air flamesare studied, as a function of the nozzle-stagnationplate separation distance, to assess the simplified hy-drodynamic model employed. Reasonable agreementis seen for all cases, provided cold-flow (upstreamof the flame) boundary conditions are specified frommeasurements. Flame-temperature effects are probedby nitrogen dilution. Results indicate that the sim-ulations can capture the hydrodynamics, as well aseffects caused by variations in flame temperature.

Flames investigated over a range of compositionsused both undiluted air-methane mixtures and dilutedair-methane mixtures to maintain an approximatelyconstant post-flame temperature rise. Along withthe variation in L/d, these preliminary experimentstarget the convective, transport, and kinetic compo-nents of the model. Reasonable agreement is seen formethane-air flames, with the results suggesting thatflame strength is slightly overpredicted, especially forlean flames. The results from these preliminary inves-tigations suggest that a complete study would providefurther insight.

Acknowledgments

We acknowledge discussions with K. Sone and as-sistance by D. Lang with digital-imaging, and by G.Katzenstein with mechanical design and assembly.The work was funded by AFOSR Grant F49620–01–1–0006, whose support is gratefully acknowledged.

References

[1] P. E. Dimotakis, F. D. Debussy, M. M. Koochesfahani,Phys. Fluids 24 (1981) 995–999.

[2] A. A. Adaczyk, L. Rimai, Exp. Fluids 6 (1988) 373–380.

[3] C. M. Vagelopoulos, An experimental and numericalstudy on the stability and propagation of laminar pre-mixed flames, Ph.D. thesis, University of SouthernCalifornia, 1999.

Page 6: Particle Streak Velocimetry

[4] T. Echekki, M. G. Mungal, Proc. Combust. Inst. 23(1990) 455–461.

[5] G. Sugiyama, Proc. Combust. Inst. 25 (1994) 601–608.[6] D. G. Goodwin, Proc. of CVD XVI and EuroCVD

Fourteen, Electrochem. Soc., 2003, pp. 155–162.[7] G. P. Smith, D. M. Golden, M. Frenklach,

N. W. Moriarty, B. Eiteneer, M. Golden-berg, C. T. Bowman, R. K. Hanson, S. Song,J. W. C. Gardiner, V. V. Lissianski, Z. Qin,http://www.me.berkeley.edu/gri mech/.

[8] R. J. Kee, M. E. Coltrin, P. Glarborg, Chemically re-acting flow - theory & practice. John Wiley & Sons,Inc., New Jersey, 2003.

[9] R. J. Kee, J. A. Miller, G. H. Evans, G. Dixon-Lewis,Proc. Combust. Inst. 22 (1988) 1479–1494.

[10] C. K. Law, C. J. Sung, G. Yu, R. L. Axelbaum, Com-bust. Flame 98 (1994) 139–154.

[11] C. J. Sung, C. K. Law, R. L. Axelbaum, Combust. Sci.Technol. 99 (1994) 119–132.

[12] C. J. Sung, J. S. Kistler, M. Nishioka, C. K. Law, Com-bust. Flame 105 (1996) 189–201.

[13] C. J. Sung, J. B. Liu, C. K. Law, Combust. Flame 106(1996) 168–183.

[14] J. C. Rolon, D. Veyante, J. P. Martin, F. Durst, Exp.Fluids 11 (1991) 313–324.

[15] C. D. Carter, J. M. Donbar, J. F. Driscoll, Appl. Phys.B 66 (1998) 129–132.

[16] J. Luque, R. J. H. Klein-Douwel, J. B. Jeffries, D. R.Crosley, Appl. Phys. B 71 (2000) 85–94.

[17] J. W. Daily, E. W. Rothe, Appl. Phys. B 68 (1999) 131–140.

[18] J. Luque, D. R. Crosley, LIFBASE: database and spec-tral simulation program (Version 1.5), Report No. MP99-009, SRI International, 1999.

[19] J. F. Grcar, R. J. Kee, M. D. Smooke, J. A. Miller, Proc.Combust. Inst. 21 (1986) 1773–1782.

[20] I. Yamaoka, H. Tsuji, Proc. Combust. Inst. 20 (1984)1883–1892.

[21] C. M. Vagelopoulos, F. N. Egolfopoulos, C. K. Law,Proc. Combust. Inst. 25 (1994) 1341–1347.

[22] D. L. Zhu, F. N. Egolfopoulos, C. K. Law, Proc. Com-bust. Inst. 22 (1988) 1537–1545.

Table 1: Experimental conditionsΦ L/d % O2 : (O2 + N2)

(a) 0.98 0.6 21.00.96 0.8 21.00.97 1.0 21.00.97 1.2 21.0

(b) 0.98 0.8 18.0(c) 0.69 0.8 21.0

0.76 0.8 21.00.87 0.8 21.01.08 0.8 18.01.20 0.8 19.51.31 0.8 21.0

Fig. 1: (a) PSV in a strained premixed methane-air flame(image cropped in the vertical dimension). Φ = 0.87,L/d = 0.8. (b) PSV processing implementation.

Fig. 2: Example CH PLIF images for a methane-air flame(344 × 260pix2,Φ = 0.96,L/d = 0.8). (a) single expo-sure. (b) average of 1000 exposures.

excitation wavelength [nm]387 388 389 390 391 392

0.0

0.3

0.5

0.8

1.0 (b)0.0

0.2

0.4

0.6

0.8

1.0 (a)

Fig. 3: Normalized experimental and simulated CH fluores-cence spectra in a methane-air flame (Φ = 0.96,L/d =0.8). (a) Measured excitation/response scan. (b) Spectrumsimulation performed using LIFBASE [18].

Page 7: Particle Streak Velocimetry

σ [s-1]

Su,

ref[c

m/s

]

0 50 100 150 200 2500

10

20

30

40(a)

Φ

Su0[c

m/s

]

0.7 0.8 0.9 10

10

20

30

40 (b)

Fig. 4: (a) Flame speeds, Su,ref , vs. strain rate, σ, for aΦ = 0.75, methane-air flame. Extrapolation to zero strainrate yields laminar flame speed estimate, S 0

u , and error bar.(b) Estimated laminar flame speeds of methane-air flames(squares). Data by Yamaoka & Tsuji (triangles) [20] andVagelopoulos et al. (diamonds) [21]. Numerical simula-tions use GRI-Mech 3.0 (line).

x [mm]

u[m

/s],

[CH

]/[C

H] m

ax

0 1 2 3 4 5 6 7 8 90.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

PSVCantera velocityBernoulli velocityCH PLIFCantera CH profile

Fig. 5: Φ = 0.98, L/d = 0.6, methane-air flame profiles.Simulation boundary at � = 5 mm.

x [mm]

u[m

/s],

[CH

]/[C

H] m

ax

0 1 2 3 4 5 6 7 8 90.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

PSVCantera velocityBernoulli velocityCH PLIFCantera CH profile

Fig. 6: Φ = 0.96, L/d = 0.8, methane-air flame profiles.Simulation boundary at � = 6 mm.

x [mm]

u[m

/s]

,[C

H]/

[CH

] max

0 1 2 3 4 5 6 7 8 90.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0PSVCantera velocityCantera velocity (Φ - 1.4%)Bernoulli velocityCH PLIFCantera CH profile

Fig. 7: Φ = 0.69, L/d = 0.8, methane-air flame profiles.Simulation boundary at � = 6 mm.

x [mm]

u[m

/s]

,[C

H]/

[CH

] max

0 1 2 3 4 5 6 7 8 90.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0PSVCantera velocityBernoulli velocityCH PLIFCantera CH profile

Fig. 8: Φ = 1.31, L/d = 0.8, methane-air flame profiles.Simulation boundary at � = 6 mm.

Page 8: Particle Streak Velocimetry

Φ

δ CH

[mm

]

0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.40.0

0.1

0.2

0.3

0.4

0.5(b)

(xC

H,P

LIF

-x C

H,s

im)/

δ CH

0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4-5-4-3-2-1012345

(a)

Fig. 9: (a) Difference in measured (xCH,PLIF ) and pre-dicted (xCH,sim) peak locations of CH, scaled by the sim-ulated CH profile thickness (δCH), as a function of equiva-lence ratio. (b) Experimental (closed) and simulated (open)CH profile thicknesses (FWHM). Experimental data is ref-erenced to the Φ = 0.96 undiluted flame. Undiluted flamesmarked with triangles; diluted flames with squares.