Top Banner

of 21

Effect of Inter Particle Interaction

Apr 14, 2018

Download

Documents

Afshin Bi
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
  • 7/29/2019 Effect of Inter Particle Interaction

    1/21

    Aerosol Science 34 (2003) 837857

    www.elsevier.com/locate/jaerosci

    Eect of inter-particle interactions on evaporation of dropletsin a linear array

    Venkat Devarakonda, Asit K. Ray

    Department of Chemical Engineering, University of Kentucky, Lexington, KY 40506-0045, USA

    Received 29 May 2002; received in revised form 14 October 2002; accepted 14 October 2002

    Abstract

    Eects of interparticle interactions on evaporation have been examined on droplets in a linear array generated

    by a modied vibrating orice aerosol generator (VOAG). The size as well as the spacing between the droplets

    was varied by modulating the frequency of the generator, and the liquid ow rate through the orice. Ethanol

    and methanol droplets of diering initial sizes and spacings were studied, and the instantaneous evaporation

    rates as well as droplet temperatures were measured using a resonance-based light scattering technique. The

    results show that at a xed dimensionless spacing between the droplets (i.e., the ratio of the spacing to

    the droplet radius, l=a), the size and temperature versus time data are highly reproducible. The measured

    evaporation rates were normalized by the evaporation rates of an isolated droplet under identical conditionsto obtain the interaction parameter () as a function of the dimensionless spacing between the droplets. The

    value of was observed to decrease with a decrease in l=a as predicted by the theoretical models, and as

    expected, found to be the same for ethanol and methanol droplets.

    ? 2003 Elsevier Ltd. All rights reserved.

    Keywords: Multi-droplet evaporation; VOAG; Optical resonance

    1. Introduction

    Various industrial and atmospheric processes, such as evaporation and growth of cloud droplets,

    spray drying and combustion, entail heat and mass transfer from assemblages of droplets. It is well

    known that interactions among member droplets of an assemblage retard heat and mass transfer rates

    between the droplets and their surrounding medium. Heat and mass transfer processes associated with

    single isolated droplets have been the subjects of extensive examination in many theoretical ( Fuchs,

    1959; Williams, 1975; Crespo & Linan, 1975; Chang & Davis, 1988; Law, 1982) and experimental

    Corresponding author. Tel.:+1-859-257-7990; fax: +1-859-323-1929.

    E-mail address: [email protected] (A.K. Ray).

    0021-8502/03/$ - see front matter? 2003 Elsevier Ltd. All rights reserved.doi:10.1016/S0021-8502(03)00065-X

    mailto:[email protected]:[email protected]
  • 7/29/2019 Effect of Inter Particle Interaction

    2/21

    838 V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857

    studies. In recent years light scattering techniques have been used on single droplets suspended in

    electrodynamic balances to obtain high precision data on evaporation of isolated droplets in controlled

    environments (Richardson, Hightower, & Pogg, 1986; Ray, Souyri, Davis, & Allen, 1991; Davis,

    1997; Tu & Ray, 2001). Due to the complicated nature of inter-droplet interactions the resultsfrom single droplet studies cannot be used directly to model processes involving assemblages of

    droplets. Therefore, a knowledge of interparticle interactions is essential for understanding a number

    of processes involving droplet sprays and clouds.

    A number of investigators have theoretically examined the eects of interparticle interactions

    on evaporation and combustion of droplets. Umemura (1994) has reviewed the analytical models

    available in the literature for predicting the extent of interactions in quasi-steady state processes,

    while Annamalai and Ryan (1992) and Annamalai (1995) have provided extensive reviews of the

    experimental and theoretical advances made in the study of interaction eects on droplet dynamics,

    heat and mass transfer, ignition and combustion of droplet arrays. Most theoretical studies are limited

    to two interacting particles (Brzustowski, Twardus, Wojcicki, & Sobiesiak, 1979; Umemura, Ogawa,& Oshima, 1981; Rajuand & Sirignano, 1990). Labowsky (1978, 1980) has developed a more general

    treatment based on the method of images for analysis of quasi-steady evaporation and burning rates

    of interacting droplets in an arbitrary array. Later, Ray and Davis (1980, 1981) and Marberry,

    Ray, and Leung (1984) have presented a rigorous treatment of heat and mass transport between an

    assemblage of particles and a surrounding medium. The results of theoretical studies show that the

    extent of interparticle interactions depends on the geometrical parameters, such as the shape and

    dimensions of the array, as well as on the size distribution of the droplets.

    A number of investigators (Nuruzzaman, Hedley, & Beer, 1971; Sangiovanni & Kesten, 1977;

    Sangiovanni & Labowsky, 1982; Silverman & Dunn-Rankin, 1994) have experimentally examined

    the eects of interactions on evaporation, combustion and ignition of linear arrays of droplets. They

    measured changes in size using photographic imaging techniques, and found that the combustion ratesof droplets in a linear array may decrease by as much as 50% compared to the rate of an isolated

    droplet. Sangiovanni and Labowsky (1982) have observed that the extent of interactions increases

    as the separation distance between droplets decreases, and applied the method of images technique

    to estimate the interaction parameter for a linear droplet array. Konig, Anders, and Frohn (1987)

    have used a more accurate technique, based on light scattering intensity distribution along rainbow

    angles, to estimate the size and temperature changes of linear arrays of droplets. Even though this

    technique is more accurate than the imaging method (i.e., size and refractive index are accurate

    to about 2% and 4%, respectively), the applicability of this technique is restricted only to cases

    where signicant changes in size and refractive index occur over a small period of time, such as in

    high-temperature combustion. Chen, Mazumder, Chang, Swindal, and Acker (1996) have investigatedthe evaporation rates of ethanol droplets in single as well as closely spaced multiple linear streams

    generated using a vibrating orice aerosol generator (VOAG). They have used a sizing technique

    based on positions of output resonances observed in inelastic scattered intensity versus wave-number

    shift spectra. Although the temperature of the droplets varied with time, they assumed a constant

    refractive index (i.e., constant temperature) in the sizing technique, and compared the evaporation

    rate of droplets in a linear array with the evaporation rate of an isolated droplet at the wet-bulb

    temperature.

    In addition to the geometrical parameters and droplet size, the evaporation rate of a droplet in

    an array depends on the temperature and the concentration of vapor in the surrounding medium.

  • 7/29/2019 Effect of Inter Particle Interaction

    3/21

    V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857 839

    In most of the prior studies on linear arrays of droplets, the heat and mass transfer processes have

    been treated as quasi-steady, even though for typical time scales involved in the studies (i.e., under

    a millisecond), the processes were unsteady (Devarakonda, Ray, Kaiser, & Schweiger, 1998). The

    time-dependent droplet temperature has not been measured or taken into consideration in the previousstudies, although light scattering techniques can be applied for in-situ measurement of time-dependent

    temperature (Vehring & Schweiger, 1992; Devarakonda & Ray, 2000) In addition to the temperature,

    scant attention was given to the environment surrounding the droplets. If the vapor molecules result-

    ing from evaporation of droplets are not removed instantaneously from the surrounding gas phase,

    the measured evaporation rates may decrease considerably due to the partial saturation. Furthermore,

    most studies examined interactions occurring at high separation distances between relatively large

    droplets ( 200 m) that undergo shape distortions at high velocities involved, thus aecting heat

    and mass transfer rates.

    Recently, we have demonstrated that fast processes associated with microdroplets can be studied

    in a linear stream of highly monodisperse droplets that traverse an identical path in a well-denedtime-invariant gas phase (Devarakonda et al., 1998; Devarakonda & Ray, 2000). We have used a

    resonance-based light scattering technique to determine accurately the size and refractive index of

    the droplets with high temporal resolution. The purpose of this study is to use the technique to

    examine the eect of interparticle interactions on heat and mass transfer rates between droplets in a

    linear array and the surrounding medium. Specically, we focus on the eects of interactions on heat

    and mass transfer rates at low separation distances between droplets. We have generated linear array

    of highly monodisperse droplets using a modied vibrating orice aerosol generator (VOAG) ( Lin,

    Eversole, & Campillo, 1990; Devarakonda et al., 1998). For the time scale involved in this study (i.e.,

    1 ms), the droplets evaporate under unsteady conditions, and thus, both the size and temperature

    change simultaneously with residence time. We have determined the size and the temperature of the

    droplets as functions of residence time from the resonances observed in elastic and Raman scattering,and calculated interaction parameters for both heat and mass transfer processes from the data. The

    interaction parameters for a linear array of droplets depend on the droplet size and the centercenter

    separation distance between successive droplets. We varied both these quantities by manipulating

    the ow rate of liquid through the orice and the orice vibration frequency used for generation of

    droplets.

    2. Theory

    The extent of interparticle interactions in a process involving momentum, heat or mass transfer ismeasured in terms of an interaction parameter, , which is dened as

    =RarrayRiso

    ; (1)

    where Rarray and Riso are, respectively, the transfer rate associated with a droplet in an array and

    with a single isolated droplet of identical size and composition exposed to identical environmental

    conditions. Even though prior experiments on droplets in linear arrays were conducted under un-

    steady conditions on relatively fast moving droplets, most investigators have expressed the interaction

  • 7/29/2019 Effect of Inter Particle Interaction

    4/21

    840 V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857

    parameter in terms of the quasi-steady evaporation rate of an isolated stationary droplet which is

    given by the following relation when the Lewis number (Le) is equal to 1.0 (Marberry et al., 1984)

    miso = 4Dgag ln

    1 + Cpg(T

    Tw)L

    : (2)

    Here Dg and L are the gas-phase diusivity and the latent heat of vaporization of the evaporating

    component, g and Cpg are the density and the heat capacity of the gas phase, a is the droplet

    radius, and T and Tw are the surrounding gas phase and the wet bulb temperatures, respectively.

    The above relation assumes that the Lewis number is equal to 1.0. When Eq. (2) is used, Eq. (1)

    may yield an interaction parameter, , value greater than unity for unsteady evaporation. To decipher

    the eect of the presence of other droplets in the neighborhood of a droplet, one should compare

    the evaporation rate of a droplet in an array with the evaporation rate of an identical isolated droplet

    under identical conditions. The experiments of the present study involve unsteady evaporation of

    ethanol and methanol droplets, and we shall compare our experimental results with predictions fromthe following model for unsteady evaporation of a droplet in an array.

    Consider a droplet in a linear stream of monodisperse droplets that are composed of compo-

    nent A. The droplet stream is exposed to a gas phase that is maintained at a temperature T.

    Initially, the droplet has a radius a0, and a temperature T0. In the formulation of the problem we

    assume that the gas phase convective processes can be described by heat and mass transfer coe-

    cients. The temperature prole in the droplet is given by the following partial dierential equation

    (Ray, Huckaby, & Shah, 1987)

    LCpL@TL

    @t=

    kL

    r2@

    @r

    r2

    @TL

    @r

    ; 06 r6 a(t); (3)

    where L; CpL, and kL are the density, the heat capacity and the thermal conductivity of the droplet,

    respectively, and TL is the droplet temperature at the radial position r, at time t. The following

    boundary conditions apply:

    @TL

    @r(0; t) = 0; (4)

    TL(a; t) = Ts(t): (5)

    An interfacial mass balance gives the droplet evaporation rate as

    2Ladadt

    =mShDgMA

    P0A(Ts)

    RTs

    pA

    RT

    ; (6)

    where Sh is the Sherwood number, MA is the molecular weight of component A; P0

    A(Ts) is the vapor

    pressure of A at the droplet surface temperature Ts, and pA is the partial pressure of A in the

    surrounding gas phase. Similarly, an interfacial energy balance yields the following relation for the

    droplet surface temperature:

    2akL@TL

    @r

    r=a

    = hNu kg(T Ts) + 2LLada

    dt; (7)

  • 7/29/2019 Effect of Inter Particle Interaction

    5/21

    V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857 841

    0 0.2 0.4 0.6 0.8 1

    Radial position,r/a

    276

    280

    284

    288

    292

    Drop

    lettemperature,TL

    (K)

    t=0.48 ms

    t=0.95 ms

    t=1.43 ms

    t=2.39 ms

    t=3.81 ms

    t=6.0 ms

    Parameter values:

    a0=10m, T0=T=298 K, Sh=4.24, Nu=3.86

    Fig. 1. Temperature distributions inside an isolated ethanol droplet at various times.

    where Nu is the Nusselt number, and kg is the gas-phase thermal conductivity. In the present study,

    the relation provided by Davis and co-workers (Zhang & Davis, 1987; Tain & Davis, 1987) is

    employed for the estimation of Sh and Nu. The interaction parameters, m and h, are used in

    Eqs. (6) and (7) to describe the reduction of gas-phase heat and mass transfer rates due to the

    presence of other particles. Eq. (6) shows that when the instantaneous evaporation rate and the

    surface temperature of a droplet are measured simultaneously in an experiment, m can be directly

    calculated from the experimental data, and then h can be estimated from the size and temperaturehistory of the droplets. In the absence of droplet surface temperature data, the measured droplet size

    versus time data must be matched with the values calculated from Eqs. (3) to (7) to obtain the

    interaction parameter, , by assuming m = h = .

    To elucidate the salient characteristics of unsteady evaporation, in Fig. 1, we have plotted tem-

    perature proles inside an isolated droplet (i.e., h = m = 1:0) at various times. The results were

    computed for an ethanol droplet with typical values involved in our experiments (i.e., a0 = 10 m,

    Sh = 4:2, and Nu = 3:9). The results show that at short residence times, the temperature prole in

    the droplet exhibits a sharp gradient near the surface. After about 4 ms, the temperature gradient

    disappears, and the droplet attains a nearly uniform temperature. Finally, the droplet temperature

  • 7/29/2019 Effect of Inter Particle Interaction

    6/21

    842 V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857

    SPEX

    1403

    MONO

    CHRO

    MATO

    R

    PMT 1

    PMT 2

    MOTORIZED VERTICALTRANSLATION STAGE

    BEAMTRAP

    DROPLETCHAIN

    CONSTANTPRESSURE

    TANK

    LIQUIDRESERVOIR

    FREQUENCYSYNTHESIZER

    ELECTRONICS

    VOAGAr-ion LASER BEAM

    E

    Dry Air

    DRY AIR

    FLOW CONTROLLER

    Fig. 2. Schematic of the experimental system for generating a linear stream of monodisperse droplets.

    reaches the steady-state wet-bulb temperature after about 6 ms. The time taken for a droplet to attain

    a steady state increases as the droplet size increases.

    Most of the experimental investigations on interactions in linear arrays were conducted on droplets

    larger than 50 m, and the time needed for such droplets to attain a steady state is about 150 ms, thatis considerably longer than the short residence times involved in the experiments. The droplet surface

    temperature decreases with time from the initial maximum value due to evaporative cooling, and as

    a result, the evaporation rate decreases with time. When the droplet attains a steady-state temperature

    (i.e., wet-bulb temperature Tw), the evaporation rate attains a constant value which can be calculated

    from Eq. (2) for Le = 1. During the unsteady period, the evaporation rate of an interacting droplet

    may exceed the steady-state evaporation rate of an isolated droplet, thus yielding an interaction

    parameter greater than unity when the rate is normalized by the steady-state evaporation rate. It

    should be noted that the surface of an isolated evaporating ethanol droplet initially may be cooler

    by as much as 3C compared to the droplet core, but inter-particle interactions signicantly retard

    heat and mass transfer rates associated with a droplet in a linear array. Thus, the dierence betweenthe core and surface temperatures of an interacting droplet is expected to be less than that of an

    isolated single droplet.

    3. Experiments

    Linear streams of highly monodisperse ethanol and methanol droplets were produced with a

    modied vibrating orice aerosol generator. A schematic of the experimental system is shown in

    Fig. 2. Devarakonda et al. (1998) and Devarakonda and Ray (2000) have described the experimental

  • 7/29/2019 Effect of Inter Particle Interaction

    7/21

    V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857 843

    system in detail. We shall recapitulate some of the salient features here. The droplet generation sys-

    tem is housed inside an environmentally controlled cylindrical chamber through which a gas stream

    of desired composition ows past the droplets that descend along the axis. The droplet stream is

    illuminated by a polarized Ar-ion laser beam at a xed position of the chamber. Light scattered elas-tically by the droplets is detected by two photo-multiplier tubes, located along polarization angles of

    = 00 (PMT1) and = 450 (PMT2), and Raman scattered light is collected into a double mono-

    chromator. The orice assembly is mounted on a programmable motorized vertical translation stage,

    and can be positioned accurately to any location from 1 to 10 mm distance from the laser line. Each

    droplet experiences the same history, and the axial distance from the generation point is uniquely

    related to the residence of the droplets in the chamber environment. Therefore, by examining droplets

    at various distances the variation of a droplet with time can be deciphered.

    The size of the droplets generated by a modied vibrating orice aerosol generator is given by

    a = 3Q

    4f1=3

    ; (8)

    where Q is the volumetric ow rate of liquid through the orice vibrating at a frequency f. The

    droplet generation system is highly stable; for example, under typical operating conditions, the in-

    stantaneous uctuations in the droplet size is about 1 part in 105, and the fractional long-term drift

    in the droplet size is about 105 per min. The main objective of the present study is to examine the

    eect of the nondimensional spacing, l=a, between the droplets on the reduction in the evaporation

    rate. The parameters that inuence the size (i.e., a) and spacing (i.e., l) between the droplets gener-

    ated by a VOAG are the liquid ow rate Q, and the vibration frequency of the orice f. The ow

    rate of liquid through an orice depends on the pressure dierence across the orice. In a modied

    VOAG, there is only a range of backing pressures over which stable droplets can be generated. In

    the present study, backing pressures of 1.65 bar (24 psi), 1.93 bar (28 psi) and 2.35 bar (34 psi)were used to generate ethanol droplets, and a backing pressure of 1.17 bar (18 psi) was used for

    methanol droplets.

    The velocity v of droplets generated by the VOAG is related to the separation distance l, between

    two successive droplets through the relation v = lf. An increase in the backing pressure has two

    eects: the droplet size (a) increases due to an increase in the ow rate (Q); and the distance

    between successive droplets increases due to an increase in the velocity of the droplets. It was

    found experimentally that the velocity of the droplets remains almost constant over a long range of

    frequencies. Since the liquid ow rate remains constant at a given backing pressure, an increase in

    the frequency results in a decrease in both the droplet size and the spacing between the droplets.

    Various combinations of frequencies and backing pressures were used to vary the nondimensionalspacing, l=a.

    In a typical experiment, the ow rate of liquid through the orice was maintained constant by

    keeping the backing pressure constant. To prevent the build-up of vapor arising from evaporation

    of droplets (i.e., to maintain pA = 0 in Eq. (6)) a dry air stream at a rate of 1000 ml=min was

    owed past the droplets. It was found that the air stream ow rate above 500 ml=min has no eect

    on the evaporation rate of the droplets, but a ow rate below 400 ml =min causes a reduction in the

    evaporation rate due to the partial saturation of the surrounding air. To examine the evaporation rate

    of droplets as a function of the residence time the droplet generator was moved in steps from 1

    to 7 mm distances from the laser beam. The velocity of the droplets was measured as a function

  • 7/29/2019 Effect of Inter Particle Interaction

    8/21

    844 V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857

    of downstream distance from the droplet generator head using a diraction technique (Devarakonda

    et al., 1998). The residence time of the droplets was estimated from velocity versus distance data,

    and varied from 90 to about 650 s as the droplet generator was moved from 1 to 7 mm. At a

    given position of the generator, the droplet size was varied by scanning the vibration frequency ofthe orice over a range of either 60 kHz (i.e., for long-frequency scans) or 20 kHz (for short scans)

    in steps of 10 Hz in every 0:10 s. Elastic scattering intensities detected by PMT1 and PMT2 and

    Raman scattering intensity at 2900 cm1 collected by the spectrometer were recorded simultaneously

    as functions of the droplet generation frequency. The Raman shift of 2900 cm1 corresponds to the

    CH bond stretch of ethanol and methanol molecules. The light scattering data from the long scans

    were used to determine the absolute droplet size and the refractive index, and the short scans were

    used to establish droplet size and refractive index changes with respect to the position where a long

    scan was recorded.

    4. Light scattering

    Scattered intensity versus frequency data were analyzed using LorenzMie theory (van de Hulst,

    1981; Kerker, 1983; Bohren & Human, 1983) to obtain droplet size and refractive index.

    Devarakonda et al. (1998) and Huckaby, Ray, and Das (1994) have described the theoretical basis

    as well as the data analysis procedure in detail. Here we provide a synopsis of the analysis proce-

    dure. Under the illumination of monochromatic light of wavelength , the intensity of light scattered

    elastically at a point by a homogeneous spherical particle of radius a, depends on the refractive

    index of the particle m, and the size parameter, which is dened as x = 2a=. In a dielectric or

    nearly dielectric spherical microparticle electromagnetic waves can propagate nearly loss-less, and at

    specic size parameters values resonances are excited due to the trapping of electromagnetic waveby almost total internal reection near the surface. At resonances the intensity of light inside the

    particle rises dramatically (i.e., the average internal intensity can exceed the incident light intensity

    by a few order of magnitudes). This gives rise to sharp peaks in elastic and inelastic (e.g., Raman)

    light scattering. The value of x, at which a resonance appears depends on m, and can be computed

    precisely from the theory. An experimental scattering intensity versus size parameter spectrum shows

    a number of resonance peaks. By identifying the observed resonances and by aligning them to com-

    puted resonances the size and refractive index of the scattering particle can be obtained with high

    precision (Huckaby et al., 1994; Devarakonda et al., 1998; Tu & Ray, 2001).

    For a given ow rate of liquid through the orice, the relation in Eq. ( 8) can be expressed in

    terms of the size parameter as follows:x = cf1=3; (9)

    where c is a constant. The variation of frequency causes the size parameter of the droplets to

    change according to the above relation, and as a result, experimental scattering intensity versus

    frequency spectra exhibit a number of resonance peaks. Fig. 3 shows typical experimental elastic

    and Raman scattering spectra obtained from ethanol droplets at 2.35 bar backing pressure. It should

    be noted that the elastic scattering spectrum from PMT1 (i.e., top spectrum) contains only transverse

    magnetic (TM) mode resonances, while the spectrum from PMT2 (i.e., middle spectrum) exhibits

    both transverse electric (TE) and TM mode resonances. In addition, the Raman scattering spectrum

  • 7/29/2019 Effect of Inter Particle Interaction

    9/21

    V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857 845

    200 210 220 230 240 250 260

    Frequency, kHz

    Scatteringintensity

    Output from PMT 1

    Output from PMT 2

    Raman scattered intensity at 2900 cm-1 shift

    Fig. 3. Raw experimental elastic and Raman scattering spectra obtained by varying droplet generation frequency.

    reveals more resonance peaks than those appear in the elastic scattering spectra. This is because low

    order (i.e., sharp peaks) resonances only appear in Raman scattering, but are suppressed in elastic

    scattering due to a variety of factors.

    An experimental intensity spectrum provides the frequencies at which resonances appear. Res-

    onating frequencies observed in the Raman scattering spectrum from a long scan are aligned with

    theoretical resonance positions (i.e., size parameter values) computed from LorenzMie theory to

    obtain the operating constant c in Eq. (9), and the refractive index m of the droplets at the position

    of the long scan. Devarakonda et al. (1998) and Devarakonda and Ray (2000) have described the

    alignment procedure in detail. In a nutshell, for an assumed value of m, each observed peak at fre-

    quency fi, is aligned with a calculated resonance at xi. We have considered all possible theoreticalresonances. The data, xi versus fi, thus obtained from the alignments are regressed to Eq. ( 9), and

    the alignment errors are calculated from the following relation:

    S(m) =

    Npti=1

    (xi cf1=3i )

    2: (10)

    The values of c and m at the global minimum of S are chosen as the best estimates. A visual match

    between the experimental spectrum obtained from PMT1 and the theoretical spectrum computed

    from Mie theory using the values of c and m obtained from the alignment procedure is used to

  • 7/29/2019 Effect of Inter Particle Interaction

    10/21

    846 V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857

    132 134 136 138 140 142

    Size parameter,x

    ElasticscatteringintensityfromPMT1

    Theory

    m= 1.3645, =92.14

    0

    Exp. data at t=0 min;c=8321

    Exp. data at t=50 min;c=8321

    Exp. data at t=100 min;c=8321

    Fig. 4. Three experimental elastic scattering spectra taken over a period of 100 min along with the theoretical scattering

    spectrum computed from the optimum parameter values obtained from the analysis of observed resonance peaks.

    validate the results. The alignment procedure permits determination of the operating constant c, and

    the refractive index m, of the droplets with precisions of 2 parts in 104.

    5. Results and discussion

    5.1. Reproducibility of droplet size

    At each backing pressure, light scattering data from the droplet stream were recorded by positioning

    the droplet generator at various distances from the laser beam. A typical experiment at a given

    pressure was conducted over a period of about 2 h. The reliability of such data requires a long-term

    stability in the droplet generator. To verify the reproducibility of droplet size we periodically repeatedlong-frequency scans at a reference position. Fig. 4 shows examples of three transverse magnetic

    (TM) mode elastic scattering spectra from ethanol droplets obtained from scans that were separated

    by a time interval of 100 min. Experimental spectra are indistinguishable from one another, indicating

    a long-term stability of the droplet generator. The analysis of the resonance peaks observed in

    spectra, using the peak alignment procedure outlined before, yields the same operating constant of

    c =8321, and the refractive index of m = 1:3645, for all three scans. In Fig. 4, we have compared the

    experimental spectra with the theoretical spectrum computed from Mie theory using the optimum

    values of c and m obtained from the analysis. The excellent match between the theoretical and

    experimental spectra validates light scattering results. Additionally, the light scattering results indicate

  • 7/29/2019 Effect of Inter Particle Interaction

    11/21

    V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857 847

    that at a given generation frequency the droplet size is highly reproducible (i.e., uctuates less than

    1 part in 104).

    5.2. Eect of backing pressure

    In this study we have altered the droplet size and the separation distance between the droplets in

    a linear array by varying the backing pressure used to generate droplets. An increase in the backing

    pressure causes an increase in the ow rate, and thus the droplet size, as can be seen from Eq. (8),

    increases at any given frequency. In addition, the separation distance between successive droplets

    increases as the velocity of the droplets increases. Since the liquid ow rate through the orice is

    proportional to the square root of the backing pressure, the operating constants (or the frequencies

    at which a given size droplets is generated) at two dierent backing pressures are related through

    the following expression that can be derived from Eqs. (8) and (9)

    c

    c0=

    f

    f0

    1=3

    =

    P

    P0

    1=6

    : (11)

    The above relation shows that the operating constant c, or the frequency f, at which a given size

    droplets appears at any backing pressure P, can be calculated from the known values of the operating

    constant c0, and the frequency f0, at a reference pressure P0. We have tested the relation with data

    obtained at various backing pressures. Fig. 5 shows TM mode elastic scattering intensity versus

    frequency spectra from ethanol droplets obtained from PMT1 at the same position of the droplet

    generator but at two dierent backing pressures of 1.65 and 1.93 bar. The spectra show that the

    resonance peaks labeled as A to G in the bottom spectrum obtained at 1.65 bar also appear in the top

    spectrum obtained from droplets generated at 1.93 bar backing pressure, but each of the resonance

    peaks shifts to a higher frequency. Each of the labeled resonances has nearly the same value of sizeparameter x (i.e., neglecting the eect of the small refractive index dierence between the droplets

    generated at two dierent backing pressures). Therefore, we can predict the frequency of a peak

    at the higher backing pressure from its observed value at the lower backing pressure. We nd that

    the frequency shift of about 20 kHz calculated from Eq. (11) for the increase of backing pressure

    from 1.65 to 1.93 bar is in excellent agreement with the frequency shift observed between the

    spectra.

    The application of the peak alignment procedure to the resonance peaks observed in the Raman

    scattering spectra yields c = 8320:8 and m = 1:3645, for the experiment at 1.65 bar, and c = 8580:5

    and m = 1:3643 for the experiment at 1.93 bar, respectively. The higher value of the operating

    constant c, implies that the size of droplets generated at a given frequency is larger at 1.93 barpressure than the size obtained at 1.65 bar pressure. The value of c = 8552 at 1.93 bar predicted

    from Eq. (11) using the value at 1.65 bar, matches closely with the value (i.e., c =8580:5) obtained

    from the analysis of resonance peaks. The dierence of about 0:30% between the values of c can

    be attributed to inaccuracies in pressure measurements.

    It should be noted that there is a small dierence in the refractive index (i.e., about 0.0002)

    between the droplets generated at two backing pressures. This is because the droplets generated at

    the higher pressure travel faster than the droplets at the lower pressure, and thus, the residence time

    (i.e., the time taken by droplets to travel from 0 to 4 mm) of the droplets generated at the higher

    pressure is shorter compared to the droplets at the lower pressure. Since the evaporation process

  • 7/29/2019 Effect of Inter Particle Interaction

    12/21

    848 V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857

    200 210 220 230 240 250 260Frequency (kHz)

    ScatteringintensityfromPMT1

    A

    A

    B

    C

    D

    E

    F

    G

    H

    I

    JL

    K

    M

    N

    B

    C

    D

    E

    F

    G

    H

    I

    J

    Pback=1.65 bar, z= 4 mm, c= 8320.8, m= 1.3645

    Pback=1.93 bar, z= 4 mm, c= 8320.8, m= 1.3643

    Fig. 5. Elastic scattering intensity versus frequency spectra from ethanol droplets generated at two dierent backing

    pressures. As the pressure increases the resonance peaks shift towards higher frequencies, indicating an increase in droplet

    size.

    is unsteady for the residence times involved in the experiments, the temperature of the droplets

    decreases with increasing residence time due to the evaporative cooling. Therefore, at a distance of

    4 mm from the generator the temperature of droplets generated at the lower pressure drops more (by

    about 0:5C) than the droplets generated at the higher pressure, and the small temperature dierence

    is reected through the dierence in the refractive index.

    5.3. Evaporation rates

    To determine evaporation rates of droplets at a given backing pressure we obtained elastic and

    Raman scattering spectra by scanning the frequency synthesizer over a 20-kHz range (i.e., short

    scans) at various positions of the droplet generator. Fig. 6 shows Raman spectra from ethanol droplets

    generated at 1.65 bar obtained by positioning the droplet generator at various distances from the laser

    beam. As the droplet generator moves away from the laser beam (i.e., as the residence time of the

  • 7/29/2019 Effect of Inter Particle Interaction

    13/21

    V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857 849

    230 235 240 245 250Frequency (kHz)

    Ramanscatteringintensityat2900cm-1s

    hift

    A

    A

    A

    A

    A

    A

    A

    z=1 mm

    z=2 mm

    z=3 mm

    z=4 mm

    z=5 mm

    z=6 mm

    z=7 mm

    Pback=1.65 bar

    Fig. 6. Raman intensity versus droplet generation frequency spectra at various positions of the droplet generator. Resonance

    positions shift to lower frequencies as the droplet generator moves away from the laser beam, indicating a decrease in

    droplet size due to evaporation.

    droplets increases), the resonance peaks shift to lower frequencies (e.g., see peaks labeled A). This

    indicates that at a given frequency, the droplet size decreases with increasing residence time due to

    evaporation. In a typical experiment, the short scans were repeated a number of times to examine

    the reproducibility of experimental data. From the observed frequency shifts in the resonance peaks,we have determined the changes in the operating constant c, and the refractive index m, with respect

    to the reference position values, using an alignment procedure described by Devarakonda et al.

    (1998) and Devarakonda and Ray (2000). The optimum values of c and m obtained from analysis

    of resonance peaks observed in a number of experiments on ethanol droplets are summarized in

    Table 1. The results show that at any position the maximum dierence among the operating constant

    values is only about 1.5, and among the refractive index values is 0.0001. The maximum dierence

    in c corresponds to a drift of about 1 A for a 11:0 m droplet. The high degree of reproducibility

    attests to the stability of the experimental conditions, in addition to the precision of the light scattering

    measurements.

  • 7/29/2019 Effect of Inter Particle Interaction

    14/21

    850 V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857

    Table 1

    Results obtained from ethanol droplets generated at 1.65 bar backing pressure

    Position Experiment 1 Experiment 2 Experiment 3 Experiment 4

    z c Refractive c Refractive c Refractive c Refractive

    (mm) (Hz)1=3 index, m (Hz)1=3 index, m (Hz)1=3 index, m (Hz)1=3 index, m

    7 8302.2 1.3656 8301.4 1.3657 8300.6 1.3657 8300.7 1.3657

    6 8306.3 1.3654 8306.3 1.3654

    5 8313.8 1.3649 8313.4 1.3650 8312.9 1.3650 8312.9 1.3650

    4 8321.7 1.3644 8320.8 1.3645 8320.3 1.3645 8320.4 1.3645

    3 8329.7 1.3639 8328.6 1.3640 8328.8 1.3640 8328.1 1.3640

    2 8337.8 1.3634 8336.2 1.3635 8336.5 1.3635 8336.5 1.3635

    1.5 8343.2 1.3631

    1 8348.4 1.3627 8347.6 1.3627 8346.9 1.3627 8347.5 1.3627

    The operating constant c, and refractive index m, values at various positions of the droplet gener-

    ator were translated into droplet size and temperature versus residence time data. It should be noted

    that the size of droplets generated at a given frequency is proportional to the operating constant c,

    and therefore, the size of droplets generated at that frequency can be calculated at various positions

    from the known value of size at a given position. The droplet size at a given backing pressure

    varies according to Eq. (8) during the frequency scans of the generator. We have translated c and

    m, versus z data with respect to a reference size of 11:0 m at z= 3 mm. We have converted the

    droplet generator position data to residence times using the diraction fringe measurement technique(Devarakonda et al., 1998). In Fig. 7, we have plotted droplet size change versus residence time

    results for ethanol and methanol droplets generated at various backing pressures. The experimental

    data are highly reproducible, that is, at a given backing pressure the maximum dierence between

    the data from two sets of experiments is less than 3%. The results show that the rate of size change

    decreases from the initial value as the time progresses, indicating the usual unsteady state evaporation

    characteristics for the residence times involved in this study. A comparison between the data from

    ethanol droplets generated at two dierent pressures reveals that over the same residence time the

    size of droplets generated at 1.93 bar pressure decreases more than droplets at 1.65 bar. This can

    be attributed to the eects of interparticle interactions. The evaporation rate of droplets generated

    at the lower pressure is retarded more because of the lower interparticle separation distance. Theresults also show that more volatile methanol droplets evaporate signicantly faster than ethanol

    droplets.

    The refractive indices of both ethanol and methanol decreases linearly by 0:0004 for each 1C

    increase in the temperature (Weast, 1985). We have used this information to convert refractive

    index values at various positions to droplet temperatures. The changes in the droplet temperature with

    respect to time are shown in Fig. 8. Again, the results are highly reproducible. The temperature values

    of one set of experiments are duplicated in another set of experiments under identical conditions.

    Temperature data exhibit characteristics similar to size data, that is, initially, the temperature changes

    rapidly, and then the rate of change decreases as the residence time increases.

  • 7/29/2019 Effect of Inter Particle Interaction

    15/21

    V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857 851

    0 100 200 300 400 500

    Residence time (s)

    -80

    -60

    -40

    -20

    0

    Dr

    opletsizechangea(nm)

    Ethanol droplets at 1.65 bar

    Ethanol droplets at 1.65 bar

    Ethanol droplets at 1.93 bar

    Ethanol droplets at 1.93 bar

    Methanol droplets at 1.17 bar

    Methanol droplets at 1.17 bar

    Polynomial fit

    Fig. 7. Droplet size change versus residence time data for ethanol and methanol droplets generated at various backing

    pressures.

    5.4. Interaction parameter

    To examine the eect of interparticle interactions on evaporation of droplets we have varied the

    dimensionless spacing (i.e., l=a) between droplets through manipulation of backing pressure and

    frequency used for droplet generation. We have used backing pressures of 1.65, 1.93 and 2.35 bar

    to generate ethanol droplets, and calculated the separation distance between two successive droplets,

    using the relation l = v=f.As mentioned before, at a given frequency an increase in the backing pressure increases the

    separation distance l, between droplets, while at a given backing pressure an increase in the frequency

    decreases the separation distance. Within each set of experiments (i.e., for a given backing pressure

    and frequency range), the spacing between two successive droplets decreases with increasing distance

    from the orice due to the unsteady state momentum transfer (i.e., drag eects). To minimize

    the eect of the variation in the separation distance, we considered only data obtained between

    4 and 7 mm distance from the orice, and in this range separation distance, l, varies less than

    10%. We shall present results on the basis of the average value of l=a in the range of distance

    considered.

  • 7/29/2019 Effect of Inter Particle Interaction

    16/21

    852 V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857

    0 100 200 300 400 500

    Residence time (s)

    -10

    -8

    -6

    -4

    -2

    0

    Dro

    plettemperatureechangeT(C)

    Ethanol droplets at 1.65 barEthanol droplets at 1.65 bar

    Ethanol droplets at 1.93 bar

    Ethanol droplets at 1.93 bar

    Methanol droplets at 1.17 bar

    Methanol droplets at 1.17 bar

    Polynomial fit

    Fig. 8. Droplet temperature change versus residence time data for ethanol and methanol droplets generated at various

    backing pressures.

    We calculated interaction parameters for the experimental droplet streams from the size and tem-

    perature versus residence time data. It should be pointed out here that the surface temperatures of the

    experimental droplets, as mentioned before, were cooler than the center temperatures because of the

    high evaporation rates involved in the experiments. The temperature variation in a droplet introduces

    a variation in the refractive index. The technique, based on Raman scattering resonances, used in

    this study estimates refractive index of a droplet that is averaged over a narrow region near the sur-

    face of the droplet (Lin & Campillo, 1995; Devarakonda & Ray, 2000). Therefore, the temperature

    obtained from the refractive index corresponds closely to the surface temperature of the droplet. We

    have used the temperatures estimated from the refractive index as the surface temperatures, Ts, in

    the following expression that was used to calculate experimental values of the interaction parameterfor mass transfer

    m =L

    ShDgMA

    da2

    dt

    RTs

    P0A(Ts): (12)

    It should be noted that the above relation has been obtained by rearranging Eq. (6) with pA=0, for

    the experimental conditions of this study. Since all the parameters on the right-hand side of Eq. ( 12)

    are either known or can be estimated, we can directly calculate m, from the experimental size

    and temperature versus residence time data. To obtain the interaction parameter, h, we numerically

    solved the partial dierential equation describing the temperature prole in a droplet (i.e., Eq. (3))

  • 7/29/2019 Effect of Inter Particle Interaction

    17/21

    V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857 853

    3 5 6

    Dimensionless spacing, l/a

    0.2

    0.3

    0.4

    0.5

    Interactionparameter,

    mo

    rh

    Experimental m

    Experimental h

    Polynomial fit

    4

    Fig. 9. Interaction parameter values obtained at various dimensionless separation distances.

    using the value of m obtained from the above procedure in the boundary condition, Eq. (6), forthe interfacial mass transfer. We varied the value of h, in the boundary condition, Eq. (7), and the

    value of h for which the overall changes in the size (i.e., a) and the surface temperature matched

    with the experimental changes was used chosen as the best estimate for the interaction parameter

    for heat transfer for the experimental data.

    The values of interaction parameters for heat and mass transfers obtained using the procedure

    outlined above are plotted in Fig. 9 for various experimental values of l=a. The results indicate that

    the interaction parameters for both heat and mass transfer are almost identical, that is, m =h =, and

    the interaction parameter, as expected, increases as the dimensionless separation distance between

    the droplets increases. We regressed experimental versus l=a data to a second-order polynomial to

    obtain the following expression:

    = 0:0611 + 0:0505

    l

    a

    + 0:0029

    l

    a

    2for 36

    l

    a6 6: (13)

    Theoretically, the interaction parameter for a given geometry is independent of the physical proper-

    ties of droplets, and this has been conrmed experimentally by Sangiovanni and Labowsky (1982).

    Therefore, the relation between and l=a, Eq. (13), obtained from experiments on ethanol droplets

    can be applied to estimate the interaction parameter of linear arrays of droplets containing any

    components. In order to validate the results obtained from the experiments on ethanol droplets, we

    have computed the evaporation rates of linear arrays of methanol droplets for m = h = = 0:27

  • 7/29/2019 Effect of Inter Particle Interaction

    18/21

    854 V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857

    0 100 200 300 400 500

    Residence time (s)

    10.90

    10.92

    10.94

    10.96

    10.98

    11.00

    11.02

    DropletRadius(m)

    284

    286

    288

    290

    292

    294

    Dropletsurfacetemperature

    Experimental radius

    Computed radius

    Experimental temperatureComputed temperature

    Fig. 10. Comparison of experimental droplet size and temperature versus residence time data for methanol droplets with the

    values predicted by a mathematical model using the interaction parameter value obtained from ethanol droplet experiments.

    obtained from Eq. (13) for the experimental separation distance of 3:5. The droplet size and sur-

    face temperature computed as functions of the residence time from the model (i.e., Eqs. (3)(7))

    are compared with experimental data in Fig. 10. The results show excellent agreement between the

    observed and computed droplet size and surface temperature versus residence time data. The max-

    imum deviation in the droplet size between the observed and computed values is less than 3 nm,

    and this is within the accuracy of the light scattering technique for size determination. The droplet

    temperatures computed from the model mostly deviate from the experimental values by about 0 :5 K,

    and this corresponds to the accuracy of our temperature determination from the refractive index

    data. The maximum deviation occurs initially because of the inconsistency between the initial and

    boundary conditions in the model. The experimental temperature values are consistently higher than

    the computed values. Since the evaporation process is unsteady, there is a sharp temperature gradient

    near the droplet surface, and the droplet core is slightly warmer than the surface. The temperature

    estimated from the refractive index provides a temperature that is averaged over a narrow regionnear the surface of the droplet, and thus, the refractive index may overestimate surface temperature

    by about 0:5 K.

    6. Conclusions

    We have examined evaporation of highly monodisperse droplets in linear arrays. Droplets were

    generated using a modied VOAG. We have varied the residence time (i.e., exposure time in the

    gas phase since generation) of the droplets by varying the distance between the droplet generator

  • 7/29/2019 Effect of Inter Particle Interaction

    19/21

    V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857 855

    and the laser beam used for light scattering. We have determined the size and refractive index of

    the droplets at various residence times using a light scattering technique based on the resonances

    observed in the elastic and Raman scattering spectra. The technique permits determination absolute

    size and refractive index and their changes with relative errors of about 2 parts and 1 part in 10 4,respectively. We have interpreted refractive index data to obtain droplet surface temperature as a

    function of the residence time. We have demonstrated that the size and temperature versus residence

    time data under given experimental conditions are highly reproducible.

    Because of the inter-droplet separation distances involved the eects of interparticle interactions

    were inherently present in the experiments. To examine the eects of interparticle interactions we

    have varied the non-dimensional spacing (i.e., l=a) between the droplets by manipulating the fre-

    quency and the backing pressure used for droplet generation. We have normalized the measured

    evaporation rates by the rates of single isolated droplets evaporating under identical conditions to

    obtain the interaction parameter, , as a function of the dimensionless spacing between the droplets.

    For the residence times involved in this study the evaporation process was transient, and thus, theevaporation rate as well as droplet temperature changed with time. We have directly estimated in-

    teraction parameters for mass transfer from the instantaneous evaporation rate as well as droplet

    temperature data obtained from the analysis of light scattering data. We have established interaction

    parameters for heat transfer by comparing the experimental droplet surface temperature versus resi-

    dence time data with values computed from an unsteady state evaporation model. The results show

    that the interaction parameter values obtained at various dimensionless spacings are identical for

    both heat and mass transfer processes, and decrease with decreasing dimensionless spacing between

    the droplets.

    To examine the validity and the applicability of the experimental results we have compared the

    size and the temperature versus residence time data obtained from methanol droplets with values

    computed from the unsteady state evaporation model using the interaction parameter value obtainedfrom ethanol droplet experiments. The experimental size and temperature data of methanol droplets

    agree well with the computed results. This validates the interaction parameter results obtained in

    this study, and conrms that the interaction parameters are independent of the nature of substances

    involved and depend only on the geometrical parameters of the droplet array.

    Acknowledgements

    The authors are grateful to the National Science Foundation for their support (grant # CTS-

    0130778).

    References

    Annamalai, K. (1995). Mechanics and combustion of droplets and sprays (3rd ed.). New York: Begell House.

    Annamalai, K., & Ryan, W. (1992). Interactive processes in gasication and combustion. Part 1: Liquid drop arrays and

    clouds. Progress in Energy and Combustion Science, 18, 221295.

    Bohren, C. F., & Human, D. R. (1983). Absorption and scattering of light by small particles. New York: Interscience.

    Brzustowski, T. A., Twardus, E. M., Wojcicki, S., & Sobiesiak, A. (1979). Interaction of two burning fuel droplets of

    arbitrary size. American Institute of Aeronautics and Astronautics Journal, 17, 12341242.

  • 7/29/2019 Effect of Inter Particle Interaction

    20/21

    856 V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857

    Chang, R., & Davis, E. J. (1988). Interfacial conditions and evaporation rates of a liquid droplet. American Institute of

    Chemical Engineers Journal, 34, 13101318.

    Chen, G., Mazumder, M. M., Chang, R. K., Swindal, J. C., & Acker, W. P. (1996). Laser diagnostics for droplet

    characterization: Application of morphology dependent resonances. Progress in Energy and Combustion Sciences, 22,163188.

    Crespo, A., & Linan, A. (1975). Unsteady eects of droplet evaporation and combustion. Combustion Science and

    Technology, 11, 918.

    Davis, E. J. (1997). A history of single aerosol levitation. Aerosol Science and Technology, 26, 212254.

    Devarakonda, V., & Ray, A. K. (2000). Determination of thermodynamic parameters from evaporation of binary

    microdroplets of volatile constituents. Journal of Colloid Interface Science, 221, 104113.

    Devarakonda, V., Ray, A. K., Kaiser, T., & Schweiger, G. (1998). Vibrating orice droplet generator for studying fast

    processes associated with microdroplets. Aerosol Science and Technology, 28, 531547.

    Fuchs, N. A. (1959). Evaporation and droplet growth in gaseous media. New York: Pergamon Press.

    Huckaby, J. L., Ray, A. K., & Das, B. (1994). Determination of size, refractive index, and dispersion of single droplets

    from wavelength-dependent scattering spectra. Applied Optics, 33, 71127125.

    Kerker, M. (1983). The scattering of light and other electromagnetic radiation . New York: Academic Press.

    Konig, G., Anders, K., & Frohn, A. (1987). A new light scattering technique to measure the diameter of periodically

    generated moving droplets. Journal of Aerosol Science, 17, 157167.

    Labowsky, M. (1978). A formalism for calculating the evaporation rates of rapidly evaporting interacting particles.

    Combustion Science and Technology, 18, 145151.

    Labowsky, M. (1980). Calculation of the burning of interacting fuel droplets. Combustion Science and Technology, 22,

    217226.

    Law, C. K. (1982). Recent advances in droplet vaporization and combustion. Progress in Energy and Combustion Science,

    8, 171201.

    Lin, H. B., & Campillo, A. J. (1995). Radial proling microdroplets using cavity enhanced Raman spectroscopy. Optics

    Letters, 20, 15891591.

    Lin, H. B., Eversole, J. D., & Campillo, A. J. (1990). Vibrating orice droplet generator for precision optical studies.

    Review of Scientic Instruments, 61, 10181023.

    Marberry, M., Ray, A. K., & Leung, K. (1984). Eect of multiple particle interactions on buring droplets. Combustionand Flame, 57, 237245.

    Nuruzzaman, A. S. M., Hedley, A. B., & Beer, J. M. (1971). Combustion of monosized droplet streams in stationary

    self-supporting ames. In 13th symposium on combustion, The Combustion Institute, Pittsburgh, PA (pp. 787799).

    Rajuand, M., & Sirignano, W. (1990). Interaction between two vaporizing droplets in an intermediate Reynolds number

    ow. Physics of Fluids, 2, 17801796.

    Ray, A. K., & Davis, E. J. (1980). Heat and mass transfer with multiple particle interactions part I: Droplet evaporation.

    Chemical Engineering Communication, 6, 6179.

    Ray, A. K., & Davis, E. J. (1981). Heat and mass transfer with multiple particle interactions part II: Surface reaction.

    Chemical Engineering Communication, 10, 81102.

    Ray, A. K., Huckaby, J. L., & Shah, T. (1987). Thermal eects of condensation on absorption of gases in growing

    droplets. Chemical Engineering Science, 42, 19551967.

    Ray, A. K., Souyri, A., Davis, E. J., & Allen, T. M. (1991). Precision of light scattering technique for measuring optical

    parameters of microspheres. Applied Optics, 30, 39743983.

    Richardson, C. B., Hightower, R. L., & Pigg, A. L. (1986). Optical measurements of the evaporation of sulfuric acid

    droplets. Applied Optics, 25, 12261229.

    Sangiovanni, J. J., & Kesten, A. S. (1977). Eect of droplet interactions on ignition in monodispersed droplet streams.

    In 16th international symposium on combustion, The Combustion Institute, Pittsburgh, PA (pp. 577592).

    Sangiovanni, J. J., & Labowsky, M. (1982). Burning times of linear fuel droplet arrays: A comparison of experiment and

    theory. Combustion and Flame, 47, 1530.

    Silverman, M. A., & Dunn-Rankin, D. (1994). Experimental investigation of a rectilinear droplet stream ame. Combustion

    Science and Technology, 100, 5773.

    Tain, D. C., & Davis, E. J. (1987). Mass transfer from an aerosol droplet at intermediate Peclet numbers. Chemical

    Engineering Communication, 55, 199210.

  • 7/29/2019 Effect of Inter Particle Interaction

    21/21

    V. Devarakonda, A.K. Ray / Aerosol Science 34 (2003) 837 857 857

    Tu, H., & Ray, A. K. (2001). Analysis of time-dependent scattering spectra for studying processes associated with

    microdroplets. Applied Optics, 40, 25222534.

    Umemura, A. (1994). Interactive droplet vaporization and combustion. Progress in Energy Combustion Science, 20,

    325372.Umemura, A., Ogawa, S., & Oshima, N. (1981). Analysis of interaction between two buring droplets. Combustion and

    Flame, 41, 4455.

    van de Hulst, H. C. (1981). Light scattering by small particles. New York: Dover.

    Vehring, R., & Schweiger, G. (1992). Optical determination of the temperature of transparent microparticles. Applied

    Spectroscopy, 46, 2527.

    Weast, R. C. (Ed.) (1985). CRC handbook of chemistry and physics. Boca Raton, FL: CRC Press.

    Williams, A. (1975). Combustion of droplets of liquid fuels: A review. Combustion and Flame, 21, 116.

    Zhang, S. H., & Davis, E. J. (1987). Mass transfer from a single microdroplet to a gas owing at low Reynolds number.

    Chemical Engineering Communication, 50, 5165.