Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Downloaded from SAE International by Purdue University, Tuesday, August 21, 2018

~. ~ The Engineering Society


~ For Advancing Mobility
~ Land Sea Air and Space® 400 COMMONWEALTH DRIVE WARRENDALE, PA 15096

872089

The Tab Method for Numerical Calculation


of Spray Droplet Breakup
Peter J. O'Rourke and Anthony A. Amsden
Theoretical Div. Group T-3
Los Alamos National Laboratory
Los Alamos, NM

International Fuels and Lubricants


Meeting and Exposition
Toronto, Ontario
November 2-5, 1987
Downloaded from SAE International by Purdue University, Tuesday, August 21, 2018

The appearance of the code at the bottom of the first page of this paper indicates
SAE's consent that copies of the paper may be made for personal or internal use, or
for the personal or internai use of specific clients. This consent is given on the con-
dition, however, that the copier pay the stated per article copy fee through the
Copyright Clearance Center, Inc., Operations Center, 21 Congress St., Salem, MA
01970 for copying beyond that permitted by Sections 107 or 108 of the U.S. Copyright
Law. This consent does not extend to other kinds of copying such as copying for
general distribution, for advertising or promotional purposes, for creating new collec-
tive works, or for resale.

Papers published prior to 1978 may also be copied at a per paper fee of $2.50 under
the above stated conditions.

SAE routinely stocks printed papers for a period of three years follOWing date of
publication. Direct your orders to SAE Order Department.

To obtain quantity reprint rates, permission to reprint a technical paper or per-


mission to use copyrighted SAE publications in other works, contact the SAE Publica-
tions Division.

All SAE paptn 1m. abJtTacttd and indued


in the SAE Global Mobfliry !hlcbIJJl:

SA£ GLOBAL MOBILITY DATABASE

No part of this pUblication may be reproduced In any form,


In an electronic retrieval system or otherwise, wllhoutthe
prior written permission of the publisher.

ISSN 0148·7191
Copyright 1987 Society of Automotive Engineers, Inc.
This paper is subject to revision. Statements and opinions Persons Wishing to submit papers to be considered for
advanced In papers or discussion are the author's and are presentation or publication through SAE should send the
his responsibility, not SAE's; however, the paper has been manuscript or a 300 word abstract of a proposed manu-
edited by SAE for uniform styling and formal. Discussion script to: Secretary, Engineering Activity Board, SAE.
will be printed with the paper if It is published In SAE
Transactions. For permission to publish this paper In full or Printed in U.S.A.
In part, contact the SAE Publication Division.
Downloaded from SAE International by Purdue University, Tuesday, August 21, 2018

872089

The Tab Method for Numerical Calculation


of Spray Droplet Breakup
Peter J. O'Rourke and Anthony A. Amsden
Theoretical Div. Group T-3
Los Alamos National Laboratory
Los Alamos, NM

ABSTRACT as in an Eulerian fluid dynamics calculation, and


derivatives of f are approximated by taking finite
We present a method for calculating drop differences ofthe cell values. This approach, which
aerodynamic breakup in engine sprays. A short we call the full spray equation method, suffers
history is first given of the major milestones in the from two principal drawbacks. First there are
development of the stochastic particle method for large numerical diffusion and dispersion errors (5)
calculating liquid fuel sprays. The most recent ad- associated with convection through the fixed
vance has been the discovery of the importance of Eulerian mesh. Second, the computer storage re-
drop breakup in engine sprays. We present a new quirements are enormous. For example, in two
method, called the TAB method, for calculating space dimensions, the distribution function f has at
drop breakup. Some theoretical properties of the least six independent variables. Since at least ten
method are derived; its numerical implementation divisions are required to resolve changes in each
in the computer program KIVA is described; and variable, at least 106 computational cells are
comparisons are presented between TAB-method required -- exceeding the storage limits of modern
calculations and experiments and calculations computers.
using another breakup model. In a second approach to solving the spray
equation, which has been used since the early
INTRODUCTION AND BACKGROUND sixties (6,7), the spray is discretized into computa-
tional particles that follow drop characteristic
Until recently, the detailed analysis of prac- paths. Each particle represents a number of drops
tical sprays was impossible due to the complexity of identical size, velocity, and temperature. Actu-
of the physical processes occurring. The first step ally, the early particle methods only calculated
toward solving this problem was taken when a sta- individual droplet trajectories, assuming the drops
tistical formulation was proposed for spray ana- had no influence on the gas. A later method (8),
lysis (1) for similar reasons that motivated most which was restricted to steady-state sprays, in-
common approaches to turbulence modeling in sin· cluded the complete coupling between the drops
gle phase fluid flows. But even with this simplifi- and gas. This later method also discretized the as-
cation, the mathematical problem was formidable sumed droplet probability distribution function at
and could be analyzed only when very restrictive the upstream boundary, which is determined by
assumptions were made. This is because the sta- the atomization process, by subdividing the do-
tistical formulation required the solution of the main of coordinates into computational cells. Then
spray equation determining the evolution of the one parcel was injected for each cell.
probability distribution function of droplet loca- In an important advance in numerical meth-
tions, sizes, velocities, and temperatures. The ods for sprays, Dukowicz (9) suggested that the
spray equation resembles the Boltzmann equation ideas of the Monte Carlo method could be combined
of gas dynamics (2) but has more independent with particle methods for spray calculations. For
variables and more complex terms on its right- example in the method of Dukowicz, which we call
hand side representing the effects of collisions, the stochastic particle method, the distribution of
breakups, and nucleations. drops at the upstream boundary is sampled sto-
Two numerical methods have been used for chastically by a relatively small number of compu-
the solution of the spray equation. In the first tational particles. The droplet distribution func-
(3,4), the full distribution function f is found ap- tion is obtained by averaging over a long time in
proximately by subdividing the domain of coor- steady-state calculations, or over many calcula-
dinates accessible to the drops -- including their tions in unsteady problems. The stochastic parti-
physical position, velocity, size, and temperature-- cle method can calculate unsteady sprays, and it
into computational cells and keeping a value offin accounts for the full coupling due to mass, momen-
each cell. The computational cells are fixed in time tum, and energy exchanges between the drops and

0148-7191/87/1102-2089$02.50
Copyright 1987 Society of Automotive Engineers, Inc.
Downloaded from SAE International by Purdue University, Tuesday, August 21, 2018

2 872089

"as. It is robust, economical, and has provided a ure of the relative importance of gas aerodynamic
framework within which to include some impor- forces that distort a drop and surface tension forces
tant new physical effects in spray calculations. In that restore sphericity. Second, the effects ofliquid
particular, using the stochastic particle method viscosity arc included. Although these effects are
much progress has been made in discoverin" the ne"ligible for lar"e drops, liquid viscosity can si,,-
mechanisms that determine spray droplet sizes. nificantly affect the oscillations of small drops.
The first major extension of the stochastic Third, the model predicts the state of oscillation
particle method was supplied by O'l~ourke (10), and distortion of droplets. Thus, if information is
who developed and applied a method for calculat- available on how distortions and oscillations affect
in" droplet collisions and coalescences. Consistent the exchange rates of muss, momentum, and cncr-
with the stochastic particle method, collisions are "y between the droplets and gas, this information
calculated by a statistical, rather than a determin- can be incorporated in the model. Fourth, the
istic, approach. 'I'he probability distributions "ov- model gives drop sizes that arc more consistent
erning the number and nature of the collisions be- with experimentally-determined mechanisms of
tween two drops are sampled stochastically. The liquid jet breakup (18,19). There is a further ad-
method was initially applied to the diesel sprays of vantage if our droplet breakup model is used as a
lliroyasu and Kadota (11), where it was found that meanS to calculate liquid jet breakup. This is that
coalcscenccs caused a seven-fold increase in the the model predicts a velocity of the product drops
mean drop size (10). Many subsequent studies (12- normal to the path of the original parent droplet.
14) have corroborated the importance of drop This normal velocity determines an initial spray
collisions in diesel-type sprays. angle that is in good agreement with measured
A second major extension of the stochastic spray angles (18). Thus, there is no need to input
particle method was the recent addition by Reitz the spray angle.
and Diwakar of a method for calculating droplet The major limitation of the TAB model is
breakup (15,16). In comparisons of calculations that we can only keep track of one oscillation
and experiments, it was found that drop breakup mode, and in reality there arc many such modes.
was important in the hollow-cone and full-cone Thus, more accurately, the Taylor analogy should
sprays typically used in direct-injected stratified be between an oscillating droplet and a sequence of
char"e en"ines. In fact, the drop sizes downstream spring-mass systems, one for each mode of oscilla-
of the injector were found to be determined prima- tion. We keep track only of the fundamental mode
rily by a competition between coalescences and corresponding to the lowest order spherical zonal
breakups. Heitz and Diwakar (16) also suggested a harmonic (20) whose axis is aligned with the rela-
numerical method for calculating atomization that tive velocity vector between droplet and gas. This
uses a droplet breakup model. Hereafter we shall is the longest-lived and, therefore, the most impor-
for brevity use the name Heitz when referring to tant mode of oscillation, but for large Weber num-
this work. In this method, one injects droplets bers other modes are certainly excited and contrib-
whose diameter equals the nozzle exit diameter. ute to drop breakup. Despite this limitation, we
The breakup of these lar"e drops is then accom- get good agreement between our theory and exper-
plished by the breakup model. This method for cal- imentally observed breakup times.
culating atomization makes the reasonable as- In the following section we give the equations
sumption that the dynamics and breakup of a liq- used by the TAB method. These equations contain
uid jet column arc indistinguishable from those of four dimensionless constants that are determined
a train of drops with equal diameter. Although it by some theoretical and experimental results. It is
requires further experimental validation, the next shown how the model predicts and continu-
method promises to remove one of the major weak- ously connects breakup times experimentally ob-
nesses of current spray calculations -- the uncer- served for the "bag" and '\itripping" breakup
tainty in the specification of upstream boundary regimes. The bag mode occurs when the Weber
conditions. number is slightly larger than a critical value, and
'I'he purpose of the present paper is to present the stripping mode occurs for Weber numbers
an alternative model for droplet breakup and to in- much larger than this same critical value.
dependently corroborate the findin"s of Heitz and We next show how the TAB model predicts
Diwakar concerning the importance of droplet the velocity of the product drops normal to that of
breakup. the parent drop and how this normal velocity is
The model is based on an analo"y, sug"ested consistent with sorne measured spray an"les (18).
by Taylor (17), between an oscillating and dis- Thus, the spray angle is automatically calculated
torting droplet and a spring-mass system. The re- by the TAB method. In contrast, in the method of
storing force ofthe spring is analo"ous to the sur- Heitz (16) the spray angle must be independently
face tension forces. The external force on the mass specified when one injects particles into the compu-
is analogous to the "as aerodynamic force. To the tational domain.
analogy we have added the damping forces due to The Taylor analogy equations do not predict
liquid viscosity. We call this model the TAB product drop sizes, and we next give the product
(Taylor Analogy )lreakup) model. The TAB model drop size equation that we use and rnotivutc this
has several advanta"es over that of Heitz. One is equation by an energy conservation argument. It
that it predicts, as pointed out by Taylor (17), that is shown that for large Weber numbers, the prod-
there is not a unique critical Weber number for uct drop sizes arc determined by a Wcber number
breakup; whether or not a droplet breaks up de- criterion.
pends on the history of its velocity relative to the The numerical implementation of the model
gas. The Weber number is a dimensionless meas- is next described. Finally, we present computa-
Downloaded from SAE International by Purdue University, Tuesday, August 21, 2018

872089 3

tional results and compare these with the experi- Cv


ments of Hiroyasu and Kadota (11) and the calcu-
--Wi!
I
Y -
lations of Reitz (16). The two models give different "V CC
drop sizes near the injector because they use differ- 1 (_ k I, (4)
+ -(v y + .) Sill (,)1
ent breakup times. Downstream the models give .. v I
d
similar results when the back-pressure is lowest,
but at higher back-pressures the TAB method where
gives larger drop sizes than Reitz's calculations
and the experiments. We give some possible rea- P
-,
11 ~r
sons for the discrepancy. We=--
g
o
THE TAB MODEL EQUATIONS
We now give the equations of the model and
tell how some of the dimensionless constants are
determined. The equation of a damped, forced har- . ely
monic oscillator is yo= -1 (0),
(/

mx = F - lex - di ,
(1)
where we take x to be the displacement of the
equator of the droplet from its equilibrium posi-
tion. In accordance with the Taylor analogy, the and
physical dependencies of the coefficients in Eq. (1)
are

1-"
d

The overdamped case, ",2 < 0, occurs only for very


small drops. The quantity We is the Weber
number.
(2) The dimensionless constants CF, Ck, and Cd
are determined by comparing with one experimen-
and tal and two theoretical results. In shock experi-
ments (21) the critical Weber number for breakup
has been found to be Wecril = 6. In these
cl 1-11.' experiments td = 00 and Yo = Yo = O. Thus from
-=C -
m d 2
Eq. (4),
Pe'
where p and Pf are the gas and liquid densities, u Cv
is the re1ative velocity between the gas and drop- y(1) = - - We (I
- cos wi) _ (5)
let, r is the droplet radius, a is the gas-liquid sur- eke'J
face tension coefficient, and Pf is the liquid viscos- The model predicts breakup if and only if y > 1,
ity. Values for the dimensionless constants CI', Ck, which occurs if and only if
and Cd will be given shortly. We assume that drop
break-up occUrs if and only if x > Cbr, where Cb is
an additional dimensionless constant. C
v
Before solving Eq. (1), we nondimensionalize 2 - - We> I
x by Cbr. Letting y = x/(Cbr) and using Eq. (2) in CkG.
Eq. (1) gives Thus the model gives the experimental result if

.. Cp Pg u 2 Cko Cdll e . CkG.


y=-----;;--3 y ---2 y ' (3) --=2We. =12. (6)
C b Pe r- Pl Per C er/t
p

with breakup occurring if and only ify > 1. This is The constant Ck is obtained by matching to the
the equation we use in our breakup model. For fundamental oscillation frequency. Lamb (op. cit.,
constant relative speed u, the solution to Eq. (3) is p. 475) gives

C, = 8. (7)
For oscillations of the fundamen tal mode, Lamb
(op. cit., p. 640) has derived
(8)
Downloaded from SAE International by Purdue University, Tuesday, August 21, 2018

4 872089

for the damping coefficient. Finally, for the funda- but their breakup time, "the time to produce a
mental mode the equator oscillates with exactly trace of mist," is defined with some uncertainty.
half the amplitude of the north and south poles Reitz (16) useS
(20). We postulate that breakup occurs if and only
if the amplitude of oscillation of the north and
south poles equals the drop radius. This criterion
gives
,
bl/
=20~ ~u P .
(12b)
g

I (9) based on an extrapolation of initial mass loss rates


C =- in the experiments of Reinecke and Waldman (23).
/J 2'
Thus there are big differences in the proportional-
and, in conjunction with Eqs. (6) and (7), ity constants used for the large Weber number
breakup time. As we shall see later the computa-
tional results are sensitive to the value of this pro-
I (10)
C =- portionality constant. Further experiments and
v 3 comparisons with experiments are needed to deter-
mine its value more precisely.
COMPARISONS WITH EXPERIMENTAL When We is close to its critical value (bag
BREAKUP TIMES breakup regime) the breakup time is determined
from Eq. (5) by ",tbll = n, or
We will now show that the model predicts,
and continuously connects, experimental breakup
times in the stripping and bag hreakup regimes.
In shock experiments (21), it has been found that
for large We the breakup times are proportional to
l
bu
= II R P/
-.
80
(We::::: We
en!
.) (13)

This is just the half-period of the fundamental


mode of oscillation. Reitz (16) uses the full period:
fi.~u .
The breakup times that the model predicts for
Pf.!
(
bu
=11
;;2
P/
-
20
these experiments are obtained from Eq. (5).
When We > > 1, the drop will break up after a PREDICTION OF NORMAL PRODUCT DROP
small fraction of its oscillation period; that is, VELOCITY AND SPRAY ANGLES
"'tbu < < 1, where tIm is the breakup time. In this
limit The TAB model also predicts a velocity of the
product drops normal to the path of the parent
., 'I drop. At the time of breakup, the equator of the
(uT
bu drop is traveling outward with velocity :k = Cbry.
COswl!m == I - -2- . It seems reasonable that the product drops will
have normal velocity
which when substituted into the right-hand side of
Eq. (5) yields (14)
2 ') where Cu is approximately unity. If liquid jet
Cp W '"hu (11) breakup is calculated by injecting drops whose
1= - - l Y e - -
Ck C11 2 radius is the injector radius, then Eq. (14) gives a
spray angle in close agreement with experimental
when y = 1. Neglecting liquid viscosity, results (18). To see this, we note that from Eq. (5)

.,
w-=ck - .
0 c,',.~, 2 (15)
3 j == C (' Wet>.) filII
P/ 'k 'b

Substituting for ",2 in Eq. (11) and solving for tbll at the time of drop breakup when We is large_ Sub-
gives stituting from Eq. (15) into Eq. (14) and using the
breakup time Eq. (12) and Eq. (10), results in

I = v'3 jT;
p
~. (large lYe) . (12a)
v - r;-
~ ~ .J ~
bu u
g tun = = C V3 (16)
2 U It 3 p'
Ranger and Nicholls (22) give I

where e is the spray angle. The experimental


result (18) is
I
b"
=8 fi:~
P IJ
g
Downloaded from SAE International by Purdue University, Tuesday, August 21, 2018

872089 5

After equating Euld and E new and using y = 1 and


o- = -v'3 ---'''-:-,-:
2" (I7)
(,)2 = 8a/per3, one obtains after some algebraic
fa n
2 3 Lid manipulation
3+ -
3.6

for hole nozzles with inlet length L and diameter d. -


r
:::::: I
8f{
+- +-
PeT ".;.-.)(6[(-5)'
-- , (22)
Equations (16) and (17) agree when Cu = 1 and r 32 20 0 ~ 120,
Lid'" 11.8. Equation (16) does not, of course, pre-
dict the dependence of the spray angle on nozzle The value of K must be determined by
Lid. It will be shown in the computational results comparisons with experimentally measured drop
section that by giving an initial oscillation to .the sizes. In our calculations we have used K = 10/3
large injected drops, the ini tial spray angle can be because this predicts a product droplet Sauter
varied, and the effects of nozzle geometry changes mean radius of6 in shock experiments (see Eq. (23)
can be included in numerical calculations. below), and this is the critical Weber number in
these experiments.
DROP SIZES AFTER BREAKUP In the bag breakup regime y '" 0 at breakup,
and Eq. (22) gives
To predict the drop sizes after breakup, we
use an equation motivated by an analysis based on r
energy conservation. The analysis is not exact but - = 7/3 ,
predicts quite plausible sizes in the limits of bag r
32
and stripping mode breakup. In this analysis we
equate the energy of the parent drop before break- In a shock experiment with We very large, one can
up to the combined energies of the product drops show Eq. (22) gives
after breakup. Before breakup, the energy of the
parent drop in its own frame of reference is the r We
sum of its minimum surface energy (4nr20) and the -=-
energy in oscillation and distortion Eos e. If all the 6
latter contribution were in the fundamental mode, or
one can show that

4n 3 ') " .) II 5 2 2 ') (18) (23)


g o:~c
= -5 P, r (x' + (o'x'i = -5 Per (j' +W y') ,

In reality there is energy in other modes and we


take Thus the Weber number based on the product drop
Sauter mean radius is 6.
For the distribution of sizes of the product
(19) drops we have used the x-squared distribution.
The breakup process will result in a distribution of
sizes because many modes will be excited by aero-
where K is the ratio of the total energy in distor- dynamic interaction with the gas. Each mode will
tion and oscillation to the energy in the fundamen- produce drops of a different size. In addition, dur-
tal mode. Thus the energy of the parent drop be· ing the breakup process there will be collisions and
fore breakup is coa1escences of the product drops, resulting in col-
lisional broadening of the size distribution. For
(20)
lack of further evidence for the form of the product
Bold = 4l1r-o
')
+ K 5 Pfr 5 (J'~." + (,)-y-)
II ') ')
. drop distribution, we have chosen a x,squared dis-
tribution because this was measured at down-
After breakup we assume the product drops stream locations in the experiments of Hiroyasu
are not distorted or oscillating. Thus the energy and Kadota (11).
after breakup is the sum of the minimum surface Equation (23) is a purely aerodynamic crite-
energies of the product drops and the kinetic rion for the product drop sizes. In contrast, to de,
energy the product drops have due to their motion termine r32 Reitz (16) uses
normal to the path of the parent drop. The first
contribution is 4nr2a r/r32, where r32 is the Sauter
mean radius of the size distribution of the product (24)
drops. Using Eq. (14), the kinetic energy of the
product drops (in the frame of reference of the
parent drop), is 1/6 nr 5pey2, where we take C" = 1. which is postulated by Nicholls (21) to determine
Hence the total energy after breakup is the boundary between the bag and stripping
breakup modes. The quantity v is the kinematic
viscosity of the gas. In introducing the drop
(21) Reynolds number Re Eq. (24) implies that at large
We, viscous stripping of droplets from the parent
drop is a dominant mechanism for parent drop
breakup. This is inconsistent with experimental
Downloaded from SAE International by Purdue University, Tuesday, August 21, 2018

6 872089

drop breakup times [cf. Eq. (12)], which do not of this assumption would necessitate a more costly
depend on Re. It is also inconsistent with experi- direct numerical integration ofEq. (3).
mentally measured spray angles (18) and intact IfWe/12 + A > 1.0, then breakup is possible.
core lengths (19), which indicate a dominantly We then calculate the breakup time tbu assuming
aerodynamic mechanism ofliquidjet breakup. that the drop oscillation is undamped for its first
period. Again this will be true for all except very
NUMERICAL IMPLEMENTATION small drops. The time tbu is the smallest root
greater than t n of the equation
In this section we describe the numerical
implementation of the TAB method in the KIVA
computer program (24). KIVA is a computer code We
- + Acosltl)(f-I
II
)+(1)1= I
(26)
for calculating two- and three-dimensional fluid 12
flows with chemical reactions and fuel sprays. where
Sprays are calculated using the stochastic pa"ticle
method. In addition to arrays specifying the parti-
cle position, velocity, size, and temperature, to im- We
y"
plement the TAB method we keep two additional 12
arrays specifying the values of y andy of each par- COS III = - - -
ticle. Equation (4) is used to update the values ofy A
andy each computational cycle as is described and
below.
For each particle we first calculate We, td,
and ",2. A value of ",2 '" occurs only for very
small drops for which distortions and oscillations
° sinql = - - -
IAwl
."
y

are negligible. Thus if ",2 '" 0, we set yn+ 1 'In + 1 = If time t + 1 = tn + At is less than thu, then no
n
= 0, where the superscript n + 1 denotes the breakup occurs this time-step, and we use Eq. (25)
advanced-time value. If ",2 > 0, we next calculate
the amplitude A of the undamped oscillation: to update y and y.
Breakup is calculated only if t n < tim S.
t n + 1. In case of breakup, the breakup size r32 and
') n
A-= ( ." - -12 ) +
We 2 (''')2
y
-w .
normal velocity V 1. are evaluated using Eq. (22)
and Eq. (14) with y evaluated at tbu. The radius of
the product drops is chosen randomly from a x-
IfWe/12 + A s.1.0, then according to Eq. (4) the square distribution with Sauter mean radius r32.
value ofy will never exceed unity and breakup will To conserve mass, the number of drops N associ-
not occur. Most particles will pass the test ated with the computational particle is adjusted
We/12 + A < 1.0, and for these we simply update y according to
andy usingEq. (4):

"' n We) N"+'=N'1 (


rn+ I
_r_) .
n 3

y
n+l
= -
We
+e 'd[( y - 12 cos wtlt
12 We also add to the particle velocity a component
with magnitude V 1. normal to its relative velocity
n We vector to the gas. The direction of this added com-
(yn + Y ~ 12 )SinWAI)
ponent is randomly chosen in a plane normal to the
+: (25a) relative velocity vector. This procedure does not
conserve momentum in detail but it does so on the
and average. Following breakup, we assume the prod-
uct drops are not distorted or oscillating, and ac-
cordinglywesetyn+1 = yn+l = 0.
We _ vll + 1) ",
( COMPUTATIONAL RESULTS
12 - 'd
j/I+I = + we
I
d The experimental results of Hiroyasu and
Kadota (ll) have often been used (9,10,16) to vali-
.n We date numerical spray models because drop sizes
12 were measured, albeit at only one axial location.
{: ( yll + _Y_d__) cos wAf In the experiment, an axisymmetric solid-cone
1 diesel spray was injected into a chamber in which
the back pressure was varied but the temperature
- (yn _ 7: )sinwAI) (25b)
was maintained at 293 K. Spray angle and tip pen-
etration were measured from photographs of the
backlighted spray. The drops were collected down-
These formulas assume that the coefficients stream in an emulsion that preserved their size.
ofEq. (3) are constant for the duration oftime-step Average sizes and size distributions were reported
Cit. This is only approximately true, but relaxation for back pressures of 1.1,3.0, and 5.0 MPa.
Downloaded from SAE International by Purdue University, Tuesday, August 21, 2018

872089 7

The comparison calculations of this report


were performed with the mesh shown in Fig. 1.
The computational region was 15 em in the axial constant
direction, which was resolved with 40 cells, and 3 pressure
em in the radial direction, which was resolved with
25 cells. The cell dimensions were expanded with
both axial and radial distance from the injector, symmetry
where the smallest cells had radial cell size or = axis
40 cells
0.05 em and axial cell size oz = 0.20 em. The left (15 cm)
boundary was a symmetry axis, the hottom bound-
ary was a rigid free-slip wall, and the right and top
boundaries were open constant pressure boundar-
ies that allowed the flow to either enter or exit the or=0.05
mesh. oz=0.20
The Reitz method (16) was used to calculate
atomization. Thus drops were injected with radius
0.015 em, equal to the nozzle exit radius. These =t:==--wall
were injected at the lower left corner of the mesh in
the axial direction. The velocity of the injected
drops was calculated assuming a nozzle discharge
coefficient of 0.705. In baseline calculations we
injected 1.7 X 105 computational particles each Fig.1. Computational mesh for the numerical
second of problem time. This gave between 500 calculations of the Hiroyasu sprays.
and 3000 particles in the computational mesh at
steady-state conditions, the number varying with
variation of the back-pressure. The results re- SPRAY TIP
ported below did not cbange appreciably when this
particle injection rate was varied. PENETRATION
With the exception of the breakup model, the A ti A II ldf'lo J lAlculllled
same version of the KJV A program was used as in (> ('I r., 30 UPI'! P~nt Model
the study of Reitz (16). In particular, drop collis- + + 50 WP"
ions and coalescences were calculated (10), and the
drop-turbulence interaction effects were included o
(16). o
o
Figure 2 shows the computed and experimen- _S!---- .-.- ._._ .-'_.
tally measured spray tip penetrations. The com-
puted tip was defined as that axial location below
5 _.- +

which 99% of the spray mass resided. Good agree-


ment was obtained, with differences between ex-
periment and calculation being comparable to the
differences obtained on successive calculations
using different particle injection rates. Reitz (16)
obtained similar agreement. Spray tip penetration
is fairly insensitive to many important physical Fig. 2. Computed and measured spray tip
parameters (10,16), and therefore this agreement, penetrations.
while gratifying, is not sufficient for model
validation.
Spray angle is another global parameter that ", '.;;' ,(.
is not very sensitive to physical parameter varia- ,
tions. Figure 3 gives the computed and measured
spray angles and shows typical spray particle plots
at steady-state conditions. The calculated angle
was defined as the smallest apex angle of a cone
that contained 99% of the spray mass and whose
apex was located at the injector. Evidently this
definition gives a much wider cone than that meas-
ured from backlighted photographs. That the ex-
perimental comparison is not that bad, can be seen
by the fact that for each back-pressure the ratio of
the calculated to experimentally-measured spray
angle is nearly 1.7,
U liP" 3.0 liP. 5.0 liP.

Hiroyosu 4.50 Hiroyosu 6.rf Hiroyasu 7.rf

Calculated 7.fP Calculated. lO.3J Calculated 11.4°

Fig.3. Computed and measured spray angles and


spray particle posi tion plots.
Downloaded from SAE International by Purdue University, Tuesday, August 21, 2018

8 872089

Figures 4-6 are plots of the computed Sauter 160


mean radi us versus axial distance from the injec-
tor, for the calculations of this paper and of Reitz ,<0
(16). Also shown on each plot is the one reported
data point ofHiroyasu and Kadota (11). Generally ,20
the curves can be broken into two sections. Close 1iJ
to the injector the drop sizes diminish rapidly as Z 100
the large injected drops break up. Further down· :Ef,l
stream the drop sizes increase gradually because of 80
drop coalescence. The most obvious difference C!1.
between the calculations occurs near the injector,
where the calculations of Reitz (16) have much '"::srn 60

larger drop sizes. This is due to the longer breakup <0


time used in the study of Reitz [cf. Eq. (12)], which
delays the breakup of the large injected drops. 20
Downstream the calculations give similar
results when the back-pressure is 1.1 MPa. At
higher back pressures, the TAB method gives 20 ~ 00 00 100 120 1~ 100
larger drop sizes than Reitz's calculations between DlSl'ANCE FROM INJECTOR (MM)
20 and 80 rom from the injector, and both methods
give larger drop sizes than the experiments 65 mm Fig. 5. Sauter mean radius versus distance from
from the injector. The differences between the the injector for the 3.0 MFa case.
computed sizes of Reitz and of the TAB method are
not surprising, considering the different formulas HlOr--:----:--:;::::;::::=:;-,..--i
used to calculate the sizes of breakup product drops LIl:END
lJ "" REITZ
,<0
[cf. Eqs. (22) and (24)]. Better agreement between a" KIVA
.... WWURED
the TAB method results and experiments could be
obtained by reducing either the breakUp times or 120
the sizes of breakup product drops. Since the 1iJ
breakup time we use is already small compared to 100
15
that recommended by others [cf. Eqs. (12a) and
(12b)], it seems most likely that drop sizes should
be reduced by increasing the value ofK in Eq. (22).
i 80

This will be explored in future calculations.


Figure 7 shows drop sizes with and without '"
::s
rn
60

drop breakup for the experiment with back· <0


pressure 1.1 MFa. In these calculations we in-
jected drops with an initial Sauter mean radius of 20
3 pm, as in the study of O'Rourke (10). With
breakup, drop sizes are reduced by approximately 0
40%. Comparison of the breakup curve of Fig. 7 0 20 ~ 00 00 100 120 l~ 100

with the curves of Fig. 4 shows that nearly the DlSl'ANCE FROM INJECTOR (MM)
same drop sizes are obtained downstream even
though different size drops are injected. In Fig. 4, Fig. 6. Sauter mean radius versus distance from
the injector for the 5.0 MFa case
.'0
160
LI:IlEIfD
. a_ RBI1'Z
. ...-.----..,.....---..,.....-----..,.....".-.,
1<0 o. KIVA
•• IIEASt1RI!Il
120

~ 100
'.0
:E
bl 60
C!1.

~ 60
2.0
<0

. _..J._. LO
20

20 40 80 80 100
DlSl'ANCE FROt.IINJEX:TOIl (1m)
120 140 160
0.0.J...-.,...._.,...~.....-
o 2
• 6 8
Z (CM)
10 12 . I.
.....-.,...-.,...-.,...-._l

Fig. 4. Sauter mean radius versus distance from Fig. 7. Computed Sauter mean radii in calcula-
the injector for the 1.1 MFa case. tions with and without breakup.
Downloaded from SAE International by Purdue University, Tuesday, August 21, 2018

872089 9

the injected drops had diameters equal to the REFERENCES


nozzle exit diameter. Thus drop breakup is proba.
bly significantly reducing downstream drop sizes 1. F. A. Williams, Phys. Fluids.L 541 (1958).
in the calculations of Fig. 4.
Finally we perform a calculation to see if by 2. E. H. Kennard, Kinetic Theory of Gases,
injecting drops with an initial oscillation, we could McGraw·Hill Book Company Inc. (New York,
influence the spray angle. Shown in Fig. 8 are par· 1938).
ticle plots from otherwise identical calculations
with and without an initial oscillation given to the 3. H. C. Gupta and F. V. Bracco, AlAA Journal
injected drops. It can be seen that with an initial l§, 10, 1053 (1978).
oscillation of dimensionless amplitude 50.0, the
computational particles are more dispersed near 4. L. C. Hasselman and C. K. Westbrook, "A
the injector. In order to influence the initial spray Theoretical Model for Fuel Injection in
angle, it was found that the dimensionless ampli· Stratified Charge Engines," SAE Paper
tude of the initial oscillation must be comparable 780138 (1978).
to the Weber number based on the nozzle radius
and injection velocity. Further calculations are 5. P. J. Roache, "Computational Fluid
needed to obtain the dependence of initial spray Dynamics," Hermosa Publishers
angle on initial oscillation amplitude. (Albuquerque, New Mexico, 1982).
6. G. L. Borman and J. H. Johnson, "Unsteady
Vaporiza tion Histories and Trajectories of
Fuel Drops Injected in Swirling Air." SAE
Paper 598C (1962).
7. S. Lambiris, L. P. Combes, and R. S. Levine,
"Stable Combustion Processes in Liquid
I Propellant Rocket Engines," Fifth
Colloquium of the Combustion and
I Propulsion Panel, AGARD, 1962.
I
8. C. T. Crowe, M. P. Sharma, and D. E. Stock,
,
, J. Fluids Engr., 325 (1977).
9. J. K. Dukowicz, J. Comput. Phys. 35, 2, 229
(1980).
10. P. J. O'Rourke, "Collective Drop Effects on
Vaporizing Liquid Sprays," Los Alamos
National Laboratory report LA·9069·T
(1981).
AMPO=O.O AMPO=50.0
11. H. Hiroyasu and T. Kadota, "Fuel Droplet
Fig. 8. Effect of an initial oscillation on the Size Distribution in Diesel Combustion
computed spray angle. Chamber," SAE Paper 740715 (1974).

CONCLUSION 12. L. Martinelli, R. D. Reitz, and F. V. Bracco,


"Comparison of Computed and Measured
A numerical method, called the TAB (Taylor Dense Spray Jets," Ninth International
Analogy Breakup) method, has been developed for Colloquium on Dynamics of Explosions and
calculating droplet aerodynamic breakup in spray Reactive Systems, Poitiers, France, 1983.
calculations using the stochastic particle meth·
od (9). The method has several significant advan· 13. A. U. Chatwani and F. V. Bracco,
tages over previous methods for calculating drop "Computation of Dense Spray Jets," ICLASS·
breakup (15,16). Numerical calculations using the 85, London, 1985.
TAB method and the Reitz method (16) for calcula·
ting atomization, confirm the findings of Reitz (16) 14. F. V. Bracco, "Modeling of Engine Sprays,"
that drop breakup is important in the diesel sprays SAE Paper 850394 (1985),
of Hiroyasu (11). Further experiments and experi·
mental comparisons are needed to refine the TAB 15. R. D. Reitz and R. Diwakar, "Effect of Drop
method and its dimensionless constants. Breakup on Fuel Spays," SAE paper 860469
(1986).
ACKNOWLEDGMENT
16. R. D. Reitz and R. Diwakar, "Structure of
This work was supported by the United States High·Pressure Fuel Sprays," SAE Paper
Department of Energy, Office of Energy Research, 870598 (1987).
Energy Conversion and Utilization Technologies
Program.
Downloaded from SAE International by Purdue University, Tuesday, August 21, 2018

10 872089

17. G. 1. Taylor, "The Shape and Acceleration ofa


Drop in a High Speed Air Stream" The
Scientific Papers of G. 1. Taylor, ed.' G. K.
Batchelor, Vol. III, University Press
Cambridge,1963. '

18. R. D. Reitz and F. V. Bracco, "On the


Dependence of Spray Angle and Other Spray
Parameters on Nozzle Design and Operating
Conditions," SAE Paper 790494 (1979).
19. B. Chehroudi, S. H. Chen, F. V. Bracco, and
Y. Onuma, "On the Intact Core of Full-Cone
Sprays," SAE Paper 850126 (1985).
20. H. Lamb, Hydrodynamics, Dover
Publications (New York, 1932).
21. J. A. Nicholls, "Stream and Droplet Breakup
by Shock Waves," NASA-SP-194, Eds. D. T.
Harrje and F. H. Reardon, 1972, p. 126-128.
22. A. A. Ranger and J. A. Nicholls, AIAA
Journal '1. 2, 28 (1969).
23. W. G. Reinecke and G. D. Waldman, "A
Study of Drop Breakup Behind Strong Shocks
with Applications to Flight," AVCO Report
AVSD-01I0-70-RR, May 1970.
24. A. A. Amsden, et a!., "Improvements and
Extensions to the KIVA Computer Program,"
Los Alamos National Laboratory report
LA-10534-MS (October 1985); A. A. Amsden
et a!., "KIVA: A Computer Program for Two:
and Three-Dimensional Fluid Flows with
Chemical Rea~tions and Fuel Sprays," Los
Alamos NatIonal Laboratory report
LA-10245-MS (February 1985).

'This paper is SUbject to revision. Statements and opinions ad- Persons wishing to submit papers to be considered for pre-
vanced in papers or discusaion are the author's and are his sentation or publication through SAB should send the manu~
responsibility, not SAB's; however, the paper has been edited script or a 300 word abstract of a proposed manuscript to:
by SAE for uoifonn styling and fonnat. Discussion will be SecretarY, Engineering Activity Board, SAE.
printed with the paper lflt is published in SAE Transaction..
For pennhsion to publish this paper in full or in part. contact Printed in U.S.A.
the SAE Publications DIvision.

You might also like