Вы находитесь на странице: 1из 10

Pergamon Chemical Enoineerin9 Science, Vol. 50, No. 18, pp.

2933-2942, 1995
Copyright © 1995 Elsevier Science Ltd
Printed in Great Britain. All rights reserved
0009-2509/95 $9.50 + 0.00
0009-2509(95)00131-X

A C O N T I N U O U S THERMODYNAMICS MODEL FOR


M U L T I C O M P O N E N T D R O P L E T VAPORIZATION

J. TAMIM
Department of Chemical Engineering, University of Ottawa, Ottawa, Ont., Canada K1N 6N5
and

W. L. H. HALLETT*
Department of Mechanical Engineering, University of Ottawa, Ottawa, Ont., Canada KIN 6N5
(First received 25 October 1994; revised manuscript received and accepted 24 March 1995

Abstract--This paper presents a model for the evaporation of droplets of multicomponent liquids in which the mixture
composition, properties and vapour-liquid equilibrium are described by the methods of continuous thermodynamics.
Transport equations for the parameters of the distribution function describing the mixture composition are derived and
solved numerically. The resulting solutions describe the changes in liquid and vapour composition with time as well as the
variation of vapour composition in space. Sample calculations are presented for distributions approximating commercial
petroleum fuels.

INTRODUCTION droplets generally possess some degree of internal


The evaporation of droplets of complex liquid fluid motion and exhibit behaviour intermediate be-
mixtures, containing hundreds or thousands of com- tween these extremes; this has been modelled either as
ponents, is frequently encountered in engineering, a vortex flow within the droplet (Tong and Sirignano,
particularly in combustors and engines burning com- 1986), or, more simply, as an increase in the effective
mercial petroleum fuels. Earlier treatments of this radial diffusion rate (Randolph et at., 1986; Talley and
problem have approximated the mixture as a small Yao, 1986; Bergeron and Hallett, 1989a). Very little
number of discrete components, the choice of which in appears in the literature on mixtures with more than
a complex mixture is always somewhat arbitrary. An two components, except for the calculations of
alternative, more appropriate for mixtures with large Newbold and Amundson (1973) for three-component
numbers of components, is to describe the mixture mixtures and those of Hallett and Ricard (1992) for
properties using the methods of continuous thermo- seven-component mixtures simulating commercial
dynamics, in which composition is represented by diesel fuels. A qualitative understanding of the behav-
a continuous probability density function. This paper iour of commercial petroleum fuels has been gained
presents a continuous thermodynamics model for the by the experiments of Chen and El-Wakil (1969) and
processes of evaporation, vapour diffusion and heat of Braide et al. (1979), but modelling of these fuels has
transfer for a single vaporizing droplet. Transport only been attempted on an empirical basis, using the
equations for the distribution function which de- ASTM distillation curve to describe the progress of
scribes the mixture composition are derived and droplet vaporization (Bardon et al., 1990; Rah et at.,
solved, relations for properties are developed, and 1986). There are therefore no general models for the
some typical results are presented. evaporation and combustion of complex mixtures.
Most work on multicomponent droplets has con- The idea of describing complex mixtures using con-
centrated on two-component mixtures. In mixtures tinuous functions rather than discrete components
with widely differing component boiling points trans- dates back at least 60 years (Katz and Brown, 1933;
port processes within the liquid can play a significant Bowman, 1949; Edmister, 1955), but most of the de-
role in controlling the liquid surface composition and velopment of the techniques of continuous thermo-
hence the progress of evaporation. Two limiting cases dynamics has been quite recent (Aris and Gavalas,
have been identified: the well-mixed droplet, for which 1966; Kehlen and R/itzsch, 1980; Briano and Glandt,
evaporation proceeds as simple fractional distillation, 1983; Cotterman et al., 1985; Shibata et at., 1987).
and the diffusion-limited case, in which molecular Most applications to date have been to petroleum
diffusion in the radial direction is the only liquid mixtures (Hoffman, 1968; Whitson, 1983; Chou and
phase transport process (Landis and Mills, 1974). Real Prausnitz, 1986; Peng et al., 1987; Willman and Teja,
1987; Cotterman and Prausnitz, 1990). The differences
between treatments of continuous thermodynamics in
tCorresponding author. the literature lie mainly in the choice of distribution
2933
2934 J. TAMIMand W. L. H. HALLETT
function (most authors use gamma or Gaussian distri- The diffusion of the mixture is described by deriving
butions), the selection of the distribution variable transport equations for the distribution function. Sub-
(component molecular weight, boiling point and car- stituting the component mol fraction yi above into
bon number are the most common), and the model for eq.(1), allowing the interval AI to become in-
vapour-liquid equilibrium (usually Raoult's law with finitesimally small, and integrating over I from 0 to
the Clausius-Clapeyron equation, although equa- oo yields a transport equation for the vaporizing sub-
tions of state have also been applied). Although widely stance as a whole:
accepted for modelling unit operations such as distil-
lation, the technique has not yet been applied to ~t (cyr) + V.(cv*yp) = V. {c3VyF + y~V(c/5)
detailed studies of transport processes (as in this pa-
per), with the exception of the derivation of some
transport equations by Gal-Or et al. (1975), or to - yFfof(l)V(cD,.(I)) dI } (6)
combustion.
where the average diffusivity D is defined by
VAPOUR PHASE TRANSPORT EQUATIONS

The physical problem dealt with here is that of = f o Dm(I)f(I)dl (7)


a single droplet which is suddenly exposed to a hot
and Din(I) is the component diffusivity as a function of
environment at Too and begins to heat up and vapor-
I. Weighting the equation by I before integrating
ize. Chemical reaction is not considered. The main
yields a transport equation for the mean 0 of the
task in developing a continuous thermodynamics
distribution:
model of this process is to describe the diffusion of the
mixture vapour into the surrounding gas. For a single
discrete component i, this is represented by ~-~(cyvO) + V'(cv*yvO) = V" fcDV(yvO) + yeOV(c~)

~ (cyi) + V "(cv*yl) = V "(cDimVyi) (1)

where c is the molar density, y the mol fraction, and


where a second average diffusivity must be defined as
v* the molar average velocity. This and subsequent
equations are written in vector form for generality,
D0 = ; ~ Dm(I)f(I)I d I . (9)
but for the droplet problem spherical symmetry is
assumed and spherical coordinates are used. Since in
a complex mixture the concentration of any indi- Finally, weighting with 12 results in
vidual component will be small, it has been assumed
that multicomponent diffusion can be approximated ~(cyrU, ') + V' (cv* yFq')
by Fick's law, with Dim being the effective diffusivity of
i in the mixture. In continuous thermodynamics the
molar concentrations of species are given by a distri- = V- fcbV(yv'I0 + yvqJV(c/~)
bution functionf(I), so that the mol fraction of a spe-
cies i is - YF I)I2V(cDm(I))d I (10)
Yi =ftI)iAIi (2)
where AI~ is the interval in I centred about the value where yet another average diffusivity must be defined
as
of I corresponding to species i. The distribution vari-
able I may be any suitable property, usually compon-
ent molecular weight or boiling point. Molecular b ~ = f ~ Dm(I)f(i)I ~dI. (11)
weight was chosen in the present work, but the trans-
port equations given here may be used with any The new variable ~ is the second moment about the
variable I. The distribution has the property that origin, defined as

f o f ( I ) d I = 1. (3) • = I)I 2dI = 0 2 + o.2 (12)

In the particular case of droplet evaporation, the


evaporating substance is present with overall mol where o.2 is the variance of the distribution. Further
fraction Yr (F = "fuel"), the remainder being air or equations can be derived for other moments of the
other noncondensable gas (A), so that distribution if desired, but these ones suffice for calcu-
lations with a two-parameter distribution.
yA = 1 - Ye (4) The eqs (6), (8) and (10) describe the variation of
and in the vapour phase the species mol fraction vapour phase composition in space and time by
becomes means of the "fuel" mol fraction yF and the distribu-
tion parameters 0 and u? (or o.2). They are general and
Yi = yrf(I)i Air (5) allow for transport property variation in space and
A continuous thermodynamics model 2935
time. The boundary conditions for the droplet prob- The last term in eq. (17) represents energy transport
lem are: due to interdiffusion of species. Although often ne-
glected, it is significant whenever the specific heat of
r = R: YF = YVR; 0 = 011; a2 = if2
the vapour evolved differs appreciably from that of
r ---, ~ : Yr = (yFO) = (yFqO = 0 (13) the ambient gas. Assuming a perfect gas, the enthalpy
gradient (V/~i) is
where subscript R denotes the droplet surface. It is
possible to reduce eqs (8) and (10) to equations for Vhi = Cei V T (19)
0 and q' alone by combining them with (6). However,
this results in more terms when the gradient of the while the component diffusion flux is
diffusion term is taken, and also necessitates imposing J* = - cD~mVyi. (20)
arbitrary finite values of 0 and ~ as boundary condi-
tions at r ~ o0. One could also, in theory, derive an Introducing the distribution function and integrating
equation for ~r2 instead of W by using ( I - 0)2 as gives the interdiffusion term as
a weighting function rather than 12, but since 0 itself
varies in space and time the resulting equations would - Z J*'Vh'=
be much more complex than those for W.
The transport equations have been written with
density c and diffusion coefficient D grouped together - J* Cpa}' VT. (21)
to reduce the effects of property variations, since for
a perfect gas the product (cD) is independent of pres-
Continuity requires that
sure and varies less rapidly with temperature than
D itself does. In looking for ways to simplify these
~" J* = 0 (22)
equations, it was noted that in each of the equations i=1
(6), (8) and (10) the last two terms are identical or
nearly identical in finite difference approximation, so that the diffusion flux of ambient gas ("air") is
and hence together make very little contribution to
the equations. This reduces the equations to J* = f o cDm(l)V(yFf(I))dl. (23)

O (CyF) + V'(cv*yr) = V ' ( c D V y v ) (14) Hence the interdiffusion term simplifies to

~(cyFO) + V'(cv*yeO) = V'(cDV(yvO)) (15)

ff--~(CyFW) + V" (cv* yF ~ ) = V" (cDV(yFW)). (16) -- Cea] cDI(I)V [yrf(l)] dl t • VT. (24)

One could further simplify the diffusion terms to the Integration requires an expression for Cp(l); the cor-
form (c/SV. Vyr) by assuming that the diffusivities are relation of Chou and Prausnitz (1986) was used,
constant in space, eliminating terms like (V(c/)). Vyr). which gives a linear variation with I:
Trial calculations showed that (V(c/))'VyF) could
amount to nearly 20% of (c/)V. Vyr) near the droplet Cp = ac + bcl (25)
surface; neglecting it had some effect on the vapour where ac and be both include the effect of temper-
phase profiles of yF and 0, but no significant effect on ature. Integration of eq. (24) then yields
the development of liquid properties. However, the
saving in computational effort is slight, and therefore
the terms were kept in the form given by eqs (14)-(16), [(ac -- CpA)CD + bcOc lD] r y e
with the diffusivities varying in space, for the present
work. + YF[(ac -- Cea)V(cf)) + bcV(cr)O)]
The energy equation, required to calculate the heat
transfer to the droplet, must also be cast in a continu-
ous thermodynamics formulation. For a mixture with
-yrfo Eac + bcl -- Cpa]f(I) V[cD(I)] dI }" V T .

discrete components, this equation is (26)


As with the diffusion equations, the last two terms can
+ e V.(ev*r) = v- vr- J*.Vh, be shown to be identical in finite difference approxi-
i=1
(17) mation and are cancelled out, giving the final form of
the energy equation:
where T is the temperature, 2 the thermal conductiv-
ity, and Cp the mixture specific heat:
CP~t(cT) + CpV'(cv* T) = V . 2 V T
~ = y~ C , ( I ) f ( I l d l + (1 - y~)C~A. (18)
+ [(ac - Cra)cD + bcOcl)]VyF. (27)
2936 J. TAMIMand W. L. H. HALLETT
LIQUID PHASE BALANCES where the final term in eq. (32) has been eliminated by
In view of the fact that most droplets in real life the fact that the integral of f i ( l ) is 1. Weighting
exhibit some internal circulation, the liquid phase will eq. (32) with I and 12, respectively, and integrating
be approximated as well-mixed. More exact yields equations for the liquid phase distribution
models of droplet internal transport processes have parameters:
established that internal circulation produces condi-
tions intermediate between the limiting cases of well- dt N(OL -- yFO)
mixed and diffusion-limited droplets (Randolph et al.,
1986; Talley and Yao, 1986; Tong and Sirignano,
+ c3 ~(yr0) + yFO ~r
0 (cD)
~
1986). However, the effects of departures from the
well-mixed state are most evident for two-component
mixtures with components of widely differing boiling
point; they are of much less importance for mixtures
- yFf:f(l)I ~(cD(l))dI]R (34)
with closely spaced component boiling points, and
hence are not expected to be significant for continu- dVL_ 3
dt cLR
FN(~PLyF~P)
ous mixtures. In further justification of this simplifica- L
+ c
bf-~(yr~F)+YF~Ff-~(cb)
tion, the literature cited shows that although the
concentrations of individual components in the va-
pour phase may be affected by the liquid phase model,
the effect on the total fuel concentration is much less. --yFf:f(I)12~(cD(l))dllR. (35)
The temperature of the droplet will also be approxim-
ated as uniform (although time-varying), in view of the Consistent with the treatment of the transport equa-
fact that heat conduction is much more rapid than tions, the last two terms in each of these equations will
species diffusion in liquids. be neglected, simplifying them to
With these assumptions, a mol balance on the
droplet gives the total molar flux N from the droplet N(1 -- YrR) = --cD 0yr
surface as Or s (36)

N= --- (cLV) (28)


A
where ct is the liquid molar density and A and V are
the droplet surface area and volume, respectively. This
dt CLR R"
may be expanded and solved for the rate of recession
of the surface: (38)
dR I(N RdcL'~ The density cL required in these equations is cal-
dt ~ + ~--~- ] (29) culated from the liquid mass density pL:
where R is the droplet radius. The molar flux N is cL = pL/OL. (39)
obtained by matching component fluxes at the sur- AS the mass densities of members of homologous
face. For a discrete component i in the liquid phase groups show little variation within the group, PL was
the rate of vaporization is assumed constant.
1 d R dxi An energy balance is required to describe the transi-
Ni = --~-~(X,Cz V) = x , N -- c L $ - ~ (30) ent heating of the droplet

where x is the liquid phase mol fraction. In the vapour dTL 3


phase this flux is given by dt - CpLCLR (q -- Nh:°) (40)

where q is the heat flux to the droplet surface by


Ni = Nyi~ -- c D i = ~ R" (31) conduction and (if present) by radiation, CeL is the
specific heat of the liquid and h:g the enthalpy of
Equating gives vapbrization. Radiation was computed using the cor-
~.~ cLR dxi relation of Hottel et al. (1955).
N(xi -- Ym) = --cDu, + -- -- (32)
R 3 dt
VAPOUR-LIQUID EQUILIBRIUM

Introducing the distribution functions in the liquid The choice of vapour-liquid equilibrium relation is
and vapour phases and integrating gives an equation constrained by the need for expressions which can be
for the total flux N: integrated in closed form with the distribution func-
tion. As in most other works on continuous thermo-
I c~yr c~ -
N(1 -- Yea) = --cD W -- yv -~(cD) dynamics, we shall use Raoult's law together with the
Clausius-Clapeyron equation. For a mixture of dis-

+ Yr I cD(I))dI ]R
(33)
crete components, Raoult's law is
y, = xi (Pv,/P) (41)
A continuous thermodynamics model 2937
where Pvi is the component vapour pressure and P the The origin 7 is the same in both phases. It can easily be
total pressure. Generalizing this by use of a liquid shown that if one phase is represented by a F distribu-
phase continuous distribution fL(l) yields the total tion, the other automatically becomes a F distribu-
vapour phase mol fraction tion.
The transport equations were solved using finite
YF f ~ fL(I) (Pv(I)/P) dI (42) difference techniques as described by Bergeron and
Hallett (1989b). The mixture properties required must
while integration with I and (I - 0)2 respectively as be evaluated by combining component property cor-
weighting functions gives the vapour phase mean and relations with the distribution function. If the neces-
variance: sary integrations are to be done analytically, one is
restricted to relatively simple forms of correlation
yF 0 = ° A (I)(Pv(I)/P)I dl (43) between the property and the distribution variable I;
more complex relations may require numerical inte-
gration. Simple linear correlations of the form of
yra 2 =
;o°f L ( I ) ( P v ( I ) / P ) ( I -- 0)2 dI (44)
eq. (25) or (46) were used for the present calculations;
the details are given in the Appendix.
The component vapour pressure is given by the
Clausius-Clapeyron equation:
Pv(I) = PAx}aexp[(sj.~/~)(1 -- Tn(I)/T)] (45) RESULTS
where sfg is the entropy of vaporization,/~ the univer- To illustrate the capabilities of the model, calcu-
sal gas constant, and Tn the boiling point. A number lations of the vaporization histories of droplets of
of studies have used the boiling point as the distribu- n-paraffin mixtures were made. As a test, the model
tion variable, but in view of the need to devise correla- was first applied to a very narrow distribution, with
tions for transport properties molecular weight was CtL = 100, flL = 0.1, ~' = 160 initially. This approxim-
selected instead. For a homologous family of com- ated pure n-dodecane with 0L = 170, a 2 = 1, and re-
pounds, the boiling point can be approximated as suits for liquid heating and progress of evaporation
a linear function of molecular weight: are compared in Fig. 1 with those from a single-com-
ponent model. The agreement is very good, and the
Tn(l) = an + bnl. (46) small discrepancies are mainly due to the fact that the
Using Trouton's Rule for ssg one then has the vapour single-component model used a more accurate va-
pressure relation in the form pour pressure correlation.
Two more realistic distributions were then exam-
Pv(l) = PAX}aexp [A (1 -- BI)] (47) ined, one representing a diesel fuel and the other
where gasoline. The distribution parameters selected for the
A = (slg/.R)(1 - a s / T ) initial liquid composition were:
B = b B / ( T - aB) (48) "diesel": ~L = 18.5, flL = 10, 7= 0
giving 0L = 185, tr2 = 1850
Further reduction of the vapour-liquid equilibrium
relations requires the introduction of the distribution "gasoline": CtL = 5.7, flL = 15, 7 = 0
function. The function selected for both gas and liquid
giving 0L = 85.5, tr2 = 1282.5.
phases in this work is the F distribution (also vari-
ously known as the Schultz or the Pearson Type III), These distributions are shown in Fig. 2, and the
often used to represent petroleum fractions (Whitson, ASTM distillation curves for these fuels, calculated by
1983; Cotterman et al., 1985; Willman and Teja, 1987; continuous thermodynamics simulation of the ASTM
Peng et al., 1987; Shibata et al., 1987): D86 distillation test, are shown in Fig. 3. The "diesel"
parameters were chosen to best reproduce data for
fl~F(ct~---~exp - (49) a commercial fuel given by Kallio et al. (1985), while
the "gasoline" lies roughly in the midrange of the
where I = 7 is the origin, ~t and fl are parameters ASTM D439 specifications for gasoline. Although
controlling the shape, and F(ct) is the gamma function. Peng et al. (1986) have questioned the use of an
The mean and variance are unbounded gamma distribution for crude petroleum,
these ASTM curves show that it can give very reason-
0 = ~fl + 7; O"2 = 0~fl2. (50)
able approximations of distillate fuels.
Substitution of this and eq. (47) into eqs (42)-(44) Figures 4 and 5 show histories of liquid phase prop-
gives simple relationships between the distribution erties for 100 I~m droplets of these fuels. The diesel fuel
parameters in the liquid and vapour phases (Cotter- must first undergo a heating transient before its tem-
man et al., 1985): perature is high enough to produce significant vapour
quantities. The gasoline, on the other hand, is suffi-
(0t - Y) (51) ciently volatile that its surface vapour pressure is
0 - 7 = 1 + ABa2L/(OL -- Y)
initially slightly higher than what can be sustained in
~r 2 = a f t ( 0 - 7)/(0L - 7)] 2. (52) quasi-steady vaporization, causing the vapour phase
2938 J. TAMIM and W. L. H. HALLETT

~0 i I I I l 1 400
n-dodecane

TL
*C
200
I cont, thermo.
-- --- single component
Ta
I00

m
m0
TB
*C
~ ' ~ " D i e s e l "

% 200

I00
/' / F 50

I I I I
2 4
t-s I
50 Io0
Fig. 1. Liquid temperature TL and mass fraction evaporated % evaporated
m/mofor a 1.5 mm diameter n-dodecane droplet evaporating
at To = 700°C. Comparison of predictions from continu- Fig. 3. ASTM D-86 distillation curves (boiling temperature
ous thermodynamics model with ctL = 100, flL = 0.l, 7 = 160 vs volume fraction distilled) for the distributions of Fig. 2.
with single pure component calculations. The points are data for an actual fuel, taken from Kallio et al.
(1985).

/
-'~osol'ne"o¢~ 5.7, /1=15,' ~=0
200
"gosoline" /
10-2 d=lOOpm Ts//
/ ; , /~Diesel'o~=8.5 iI /
X I\ ,=,o // e7
,,,, o
/ /'-/

10-3 .2" /

I00 ' p p/ J'llII


II / ./ mO
10-4 z
0 200
l" (=tool.weight)
400
iiII / ,/
Fig. 2. Distribution functions selected for simulation of
gasoline and diesel fuel.

/ .........
fuel concentration to decrease after the start of evap-
oration. Both fuels then enter a sort of quasi-steady 0 20 40 60
state in which the liquid temperature and mean mo- t - me
lecular weight climb nearly linearly as light compo- Fig. 4. Time histories of liquid phase properties and of va-
nents are distilled out, while the vapour phase concen- pour phase tool fraction at the surface y ~ for a 100 #m
"gasoline" droplet evaporating at T® = 700°C. The mass
tration remains nearly constant. The liquid temper-
fraction evaporated m/moand the mol fraction YR~ are in
ature remains well below the bubble point, so that percent, while T is in °C.
vapour mol fractions at the surface are fairly low; this
is quite different from the behaviour of a pure fuel,
which typically approaches the boiling point closely
(cf. Fig. 1). The reason for this is the continued transi- perature to turn up more sharply. The width of the
ent heating of the mixture droplet, which absorbs distribution, as represented by the variance, drops
some of the energy which would otherwise be avail- slowly during the process as the light end of the
able for evaporation. Near the end of the droplet distribution is removed.
lifetime only heavy components remain in the liquid, Figure 6 shows the behaviour of a much larger
causing the mean molecular weight and droplet tem- droplet of diesel. The mass flux from the surface is
A continuous thermodynamics model 2939
I I I I
"gasoline" d-IOOpm
"Diesel" /
300 l'~ i ~1 / ~i t :30ms
d: I00 "lB. /
I00 "" "" \ ,,,.e
- / ~/T//-
\
TO, ---7" /
TU / / ~ ~eC 80 \ 4o
OL / / ~ -" ..¢"g~ ~" T.IO-I \ _N T
.c.
200 = / . . - IOO o
.'m m
2O
• YFR,

I00 50

:---., .~_... YE~_ 0 2 I0 20 I00 200


r/R
Fig. 7. Vapour phase properties as a function of dimension-
O 20 40 60 less radius (r/R) for a 100 tim "gasoline" droplet evaporating
t - ms at Too = 700°C, shown at time t = 30 ms. The fuel mol frac-
Fig. 5. Time histories of liquid phase properties and of va- tion YFis in percent, while T is in °C. The calculation domain
pour phase mol fraction at the surface YER for a 100 pln extends to (r/R) = 500.
"diesel fuel" droplet evaporating at Too = 700°C. The mass
fraction evaporated m/mo and the mol fraction YFR are in
percent, while T is in °C.
with a somewhat greater yield of light fractions, re-
flected in a slightly higher value of 0L for a given
l i i / fraction evaporated; because of this, the sharp upturn
/ "Di.I" / in 0L and TL evident for the smaller droplets is absent.
/ d-l.Smm TB/ / Apart from this, no significant changes in droplet
behaviour are to be seen.
Figure 7 gives a sample of the information provided
r,./ //-.,,-._"" by the model in the vapour phase. The mean molecu-
lar weight drops with increasing radius, reflecting
primarily the increase in 0L with time. A curious
I
. /, " / - " feature is the slight "hump" in the profile of a 2, which
rises to values greater than any recorded at the sur-
face. Physically this corresponds to a broadening of
the distribution as it diffuses outwards from the sur-
face, which occurs because distributions at different

,oo.F/../d. radii can mix with each other by means of diffusion,


a phenomenon akin to dispersion in chemical reac-

? tors. This effect disappears far from the droplet, where


fuel concentrations drop to negligible levels. It is of
interest to note that this " h u m p " remained even in
predictions made with constant diffusivity, or with all
0 2 4 6 three diffusivities set equal. It is evident that the va-
t-s
pour composition varies significantly in space, a fact
Fig. 6. Time histories of liquid phase properties and of va-
which may have interesting consequences for ignition
pour phase mol fraction at the surface YFR for a 1.5 mm
"diesel fuel" droplet evaporating at To = 700°C. The mass or combustion.
fraction evaporated m/mo and the mol fraction YFR are in The computational effort required for this model is
percent, while T is in °C. essentially that for iterative solution of the energy
equation (27) and the three diffusion equations
(14)-(16) at each time step. A discrete component or
pseudocomponent model by comparison will require
roughly proportional to 1/R, so that the evaporating the energy equation plus one diffusion equation for
liquid has less cooling effect; the liquid temperature each component, the iteration and convergence pro-
now approaches more closely to the bubble point, but cesses being much the same for either model. As
still not as closely as for the equivalent single-com- substantially more than three discrete components
ponent droplet (Fig. 1). The higher liquid temperature are usually required to properly represent something
causes the droplet to follow a vaporization history as complex as a commercial fuel [for example, Hallett
2940 J. TAMIMand W. L. H. HALLETT
and Ricard (1992) needed seven components to ap- R droplet radius (m)
proximate a diesel fuel], the continuous thermo- universal gas constant, 8.314 (kJ/kmol K)
dynamics model will require considerably less com- syo entropy of vaporization (kJ/kmol K)
puter time. t time (s)
T temperature (K)
CONCLUSIONS I)* molar average velocity (m/s)
It has been demonstrated that continuous thermo- V volume (m 3)
dynamics provides a practical way of modelling the X mol fraction in liquid phase
evaporation of droplets of complex mixtures. The Y mol fraction in vapour phase
sample calculations show that a model of this sort can
be the basis for realistic representations of the evapor- Greek letters
ation and combustion behaviour of commercial pe- ct, fl, y parameters of distribution function
troleum fuels. F(z) gamma function of argument z
The main purpose of this paper has been to develop 0 mean of distribution ( = mean molecular
the basic transport equations required and demon- weight) (kg/kmol)
strate the potential of this model. For more realistic 2 thermal conductivity (W/m K)
models of commercial petroleum mixtures it will be P mass density (kg/m 3)
necessary to account for the presence of different a2 variance of distribution
families of compounds (for example cycloparaffins ~F second moment about origin of distribution
and aromatics); this can be done using multiple distri- ( = ~r2 + 0 2)
bution functions and/or by adding discrete compo-
nents, extensions which have been demonstrated in Subscripts
the literature (Cotterman et al., 1985). A rigorous A ambient non-condensable gas ( = "air")
experimental test of the calculations presented here ATM atmospheric
would require the creation of liquid mixtures whose B boiling point, bubble point
composition is known in sufficient detail to allow cr critical state
accurate construction of the necessary distribution i species i
functions. There are no data available which meet L liquid phase
these requirements, the experiments in the literature o initial value
on the evaporation or combustion of droplets of mix- R at droplet surface
tures having been performed with fuels of unspecified v vapour
composition or boiling behaviour (Braide et al., 1979; oo ambient
Chen and E1 Wakil, 1969).

Acknowledgements--The authors are grateful for the finan-


cial support of the Natural Sciences and Engineering REFERENCES
Research Council of Canada. Thanks are due to Dr. M. Abramovitz, M. and Stegun, I. A., 1970, Handbook of Math-
Margerum, now with Shell Canada Ltd., for first bringing ematical Functions. Dover, New York.
continuous thermodynamics to our attention. Aris, R. and Gavalas, G. R., 1966, On the theory of reactions
in continuous mixtures. Phil. Trans. R. Soc. A 260,
351-393.
NOTATION Bardon, M. F., Gauthier, J. E. D. and Rao, V. K., 1990,
Flame propagation through sprays of multicomponent
ac, bc vapour phase specific heat parameters, eq. fuel. J. Inst. Energy 63, 53-60.
(25) Bergeron, C. A. and Hallett, W. L. H., 1989a, Autoignition of
A surface area (m2); parameter in eq. (47) single droplets of two-component liquid fuels. Combust.
Sci. Technol. 65, 181-194.
B parameter in eq. (47) Bergeron, C. A. and Hallett, W. L. H., 1989b, Ignition char-
c molar density (kmol/m 3) acteristics of liquid hydrocarbon fuels as single droplets.
Cp specific heat (kJ/kmol K) Can. J. chem. Engng 67, 142-149.
D,D,D averaged diffusivities (eqs 7, 9, 11) (m2/s) Bowman, J. R., 1949, Distillation of an indefinite number of
Dm effective diffusivity of component in mixture components. Ind. Engng Chem. 41, 2004-2007.
Braide, K. M., Isles, G. L., Jordan, J. B. and Williams, A.,
(m2/s) 1979, The combustion of single droplets of some fuel oils
f(I) distribution function and alternative liquid fuel combinations. J. Inst. Energy
-h partial molal enthalpy (kJ/kmoi K) 52, 115-124.
hlg enthalpy of vaporization (kJ/kmol K) Briano, J. G. and Glandt, E. D., 1983, Molecular thermo-
dynamics of continuous mixtures. Fluid Phase Equilibria
I distribution variable ( = molecular weight) 14, 91-102.
(kg/kmol) Chen, C. S. and El-Wakil, M. M., 1969, Experimental and
m mass (kg) theoretical studies of burning drops of hydrocarbon mix-
N molar flux (kmol/m 2 s) tures. Proc. Instn mech. Engrs 184, Pt 3J, 95-108.
P pressure (kPa) Chou, G. F. and Prausnitz, J. M., 1986, Adiabatic flash
calculations for continuous or semicontinuous mixtures
q heat flux (W/m 2) using an equation of state. Fluid Phase Equilibria 30,
r radial coordinate (m) 75-82.
A continuous thermodynamics model 2941

Cotterman, R. L., Bender, R. and Prausnitz, J. M., 1985, APPENDIX: PROPERTY CORRELATIONS
Phase equilibria for mixtures containing very many com- Mixture properties for the vapour and liquid phases must
ponents. Development and application of continuous be generated by integrating property correlations for indi-
thermodynamics for chemical process design. Ind. Engng vidual pure components with the distribution function. In
Chem. Process Des. Dev. 24, 194-203. this work, a linear variation of each property with molecular
Cotterman, R. L. and Prausnitz, J. M., 1990, Application of weight was assumed, any variation of the property with
continuous thermodynamics to natural-gas mixtures. Rev. temperature being accounted for in the coefficients. The
Inst. Franqais P~t. 45, 633~a43. following are the correlations used. For the component dif-
Edmister, W. C., 1955, Improved integral technique for fusion coefficient in the gas phase D(I)
petroleum distillation calculations. Ind. Engng Chem. 47,
1685-1690. D(I) = (ao + boI)Oo (A1)
Gal-Or, B., Cullinan, H. T. and Galli, R., 1975, New thermo-
with factor @t~ describing the temperature variation. This
dynamic transport theory for systems with continuous
gives
component density distributions. Chem. Engng Sci. 30,
1085-1092. 1) = (ao + boO)OD
Hallett, W. L. H. and Ricard, M. A., 1992, Calculations of the
auto-ignition of liquid hydrocarbon mixtures as single I) = (ao + boW~O)Oo
droplets. Fuel 71, 225-229.
Hoffman, E. J., 1968, Flash calculations for petroleum frac- D = ( a o +--~[ysa
bD 3 + 30a 2 + 0~])OD (A2)
tions. Chem. Engng Sci. 23, 957-964.
Hottel, H. C., Williams, G. C. and Simpson, H. C., 1955, where 7s is the skewness coefficient of the distribution, which
Combustion of fuel droplets. Fifth Symposium (Interna- for a gamma distribution is given by (Abramovitz and
tional) on Combustion, The Combustion Institute, pp. Stegun, 1970):
101-129.
Hubbard. G. L., Denny, V. E. and Mills, A. F., 1975, Droplet ~s = 2 / , ~ (A3)
evaporation: effects of transients and variable properties.
Int. J. Heat Mass Transfer 18, 1003-1008. The temperature dependence was fitted with a Sutherland-
Kallio,. N. N., Moyes, B. W., Webster, G. D. and Whyte, type equation, giving
R. B., 1985, SAE Technical Paper 850052.
T5/2
Katz, D. L. and Brown, G. G., 1933, Vapor pressure and
vaporization of petroleum fractions. Ind. Engng Chem. 25, @o - b® + T" (A4)
1373-1384.
Kehlen, H. and R/itzsch, M. T., 1980, Continuous thermo- The values of the three D's proved to be close to one another,
dynamics of multicomponent mixtures. Proceedings of the but since their evaluation is simple they were not assumed
6th International Conference on Thermodynamics. equal.
Merseburg, pp. 41-51. The thermal conductivity of the vapour is likewise written
Landis, R. B. and Mills, A. F., 1974, Effect of internal dif- as
fusional resistance on the evaporation of binary droplets. 2(1) = ax + bxl. (A5)
Fifth International Heat Transfer Conference, pp. 345-349.
Newbold, F. R. and Amundson, N. R., 1973, A model for The vapour mixture conductivity was approximated as that
evaporation of a multicomponent droplet. A.I.Ch.E.J. 19, of the component having a molecular weight equal to 0, and
22-30. this value combined with the conductivity of the air using the
Peng, D.-Y., Wu, R. S. and Batycky, J. P., 1987, Application Mason-Saxena rule (Reid et al., 1987) to give the overall
of continuous thermodynamics to oil reservoir fluid sys- mixture conductivity. Attempts to use combining rules from
tems using an equation of state. AOSTRA J. Res. 3, the literature to produce a more accurate equation for the
113-122. vapour conductivity led to expressions that could not be
Rah, S.-C., Sarofim, A. F. and Be6r, J. M., 1986, Ignition and integrated analytically. The approach used here can be justi-
combustion of liquid fuel droplets. Part II. Ignition stud- fied on the basis that conductivity is only a weak function of
ies. Combust. Sci. Technol. 49, 169-184. molecular weight and that air is usually the dominant species
Randolph, A. L., Makino, A. and Law, C. K., 1986, Liquid- in the vapour phase. The temperature variation of vapour
phase diffusional resistance in multicomponent droplet conductivity was approximated by a linear function:
gasification. Twenty-first Symposium (International) on
Combustion, The Combustion Institute, pp. 601-608. a x = aKTT+ arc
Reid, R. C., Prausnitz, J. M. and Poling, B. E., 1987, The bx = bxr T + bKc (A6)
Properties of Gases and Liquids, fourth edition. McGraw-
Hilt, New York.
The vapour specific heat was taken from the correlations
Shibata, S. K., Sandier, S. I. and Behrens, R. A., 1987, Phase
of Chou and Prausnitz (1986), which gives a linear variation
equilibrium calculations for continuous and semi-continu-
with molecular weight and a cubic polynomial dependence
ous mixtures. Chem. Engng Sci. 42, 1977-1988.
on temperature. For the liquid, the specific heat per unit
Talley, D. G. and Yao, S. C., 1986, A semi-empirical ap-
mass is nearly independent of molecular weight, while the
proach to thermal and composition transients inside va-
temperature variation can be approximated by a second
porizing fuel droplets. Twenty-first Symposium (Interna- degree polynomial:
tional) on Combustion, The Combustion Institute, pp.
609-616.
CpL = "0L [aL + bLT + CLT2]. (A7)
Tong, A. Y. and Sirignano, W. A., 1986, Multicomponent
droplet vaporization in a high temperature gas. Combust.
The effective enthalpy of vaporization of a mixture h,~ is
Flame 66, 221-235.
the sum of those for the pure components weighted by their
Whitson, C. H., 1983, Characterizing hydrocarbon plus frac-
fluxes (not by their concentrations). For discrete components
tions. Soc. Pet. Engng J. 23, 683-694.
Willman, B. and Teja, A. S., 1987, Prediction of dewpoints of
semicontinuous natural gas and petroleum mixtures. 1. Nhya = ~ Nihfai. (A8)
i=1
Characterization by use of an effective carbon number and
ideal solution prediction. Ind. Engng Chem. Res. 26, Combining this with the vapour phase flux expression from
948-952. eq. (31) for Ni and introducing the distribution function for
2942 J. TAMIM and W. L. H. HALLETT

a continuous mixture gives are correlated as


Tar(I) =acr "4" bad (A13)
Nh:g = - f : cD(l)hlo(l)~ (YFf(l)),RdI
then for the mixture

+ yFf(I)IR Nhsg(I) dl. (A9) Tcr =a¢r + bcrO. (A14)


For n-paraffins, the following constants for the equations
Since h:o per unit mass at the normal boiling point is nearly above were determined by linear regression of property data
constant a m o n g members of a homologous chemical group, for the range of c o m p o u n d s from n-octane to n-hexadecane:
a linear variation of the molar hzo is a very good approxima- Vapour pressure: a s = 241.4; b8 = 1.45; sya = 87.9 kJ/
tion: kmol K (for boiling point in K).
Vapour phase diffusivity (units m2/s): ao = 2.89 x 10-9;
hfg(I) = Jan + bnl]On (A10) bo = - 6.60x 10-t2; bo = 250.
where factor On represents the temperature dependence. Vapour thermal conductivity (units W / m K ) : axr
Introducing this into the previous equation, integrating, and = 1.09 x 10-*; aKc = -- 2.37 X 10-2; bxr = - 1.91
neglecting all terms involving gradients of diffusivity yields x 10-s; bxc = 3.47 x 10 -s.
Enthalpy of vaporization (units J/kmol): aH =
2.07 X 107; bn = 1-35"105; (Tcr - TB)°'38 = 6.959.
f b 0 bncD 0 . . . . ) (All)
hfa = *n~aH + HYr I• - - - - - ~ r ( Y F e ) l a f . Critical temperature: a¢, = 440.8; bet = 1.21.
Liquid specific heat (units kJ/kmolK): aL = 2.26;
bL = -- 2.94x 10-3; CL = 9.46× 10 -6 .
The temperature dependence is based on the often-used
Watson equation (Reid et al., 1987):
These correlations work well for n-paraffins, but equations
h:o ( Tar- T ~ °'38 (A12) of this sort m a y be more difficult to develop for other
0 " = h-~,, = \ r o T - r . / homologous groups (for example aromatics) which exhibit
less similarity between members than do n-paraffins.
where subscript B refers to the normal boiling point. The Apart from the diffusivities, which were allowed to vary,
term (Tar - TB)°'38 is approximated as constant for members all vapour properties were assumed constant in space and
of a homologous group of compounds, while the critical evaluated at a reference state given by the prescription of
temperature is taken as the pseudo-critical temperature cal- Hubbard et al. (1975). All properties were updated with each
culated using Kay's rule. If component critical temperatures time step.

Вам также может понравиться