Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Available online at www.sciencedirect.

com

Proceedings of the Combustion Institute 37 (2019) 3237–3244


www.elsevier.com/locate/proci

Experimental investigation of spray characteristics of a


liquid jet in a turbulent subsonic gaseous crossflow
Mohsen Broumand, Mahmoud M.A. Ahmed, Madjid Birouk∗
Department of Mechanical Engineering, University of Manitoba, Winnipeg, Manitoba, R3T 5V6 Canada

Received 30 November 2017; accepted 3 August 2018


Available online 23 August 2018

Abstract

Two perforated plates with different solidity ratios, S=50% and 67%, were used to investigate the effect of
the velocity fluctuations of a subsonic gaseous crossflow on the spray characteristics of a liquid jet includ-
ing droplet size and velocity distributions. The experiments were conducted over a range of jet-to-crossflow
momentum flux ratio of q=16.5-172, and two gas Weber numbers of Weg =2.7 and 5.9, corresponding to
the enhanced capillary breakup and bag breakup regimes, respectively. The experimental results of this study
revealed that the distribution of droplets size associated with a turbulent and a uniform crossflow for each
specific breakup regime were approximately identical. The bimodal and single peak distributions of droplets
size, respectively, associated with enhanced capillary and bag breakup regimes were generally consistent with
the literature reports. However, the transition of the liquid primary breakup regime from enhanced capillary
to bag breakup mode was delayed in a turbulent crossflow compared to its uniform counterpart. The gen-
eral behavior of droplets size-velocity profiles were also consistent with the literature reports. Nonetheless,
complex variations in the distribution of droplets velocity when changing the crossflow turbulence intensity
were observed and linked with the presence of instabilities on the liquid jet’s surface. Finally, the present
experiments allowed shedding more light on the reason why the breakup mechanisms of a liquid jet in a
conventional uniform crossflow should not be generalized to predict the distinct breakup process of a liquid
jet in a turbulent crossflow.
© 2018 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

Keywords: Liquid jet; Spray; Turbulent crossflow; Droplet size; Droplet velocity

1. Introduction crossflow is one of the most efficient spray genera-


tion techniques when rapid mixing is desired [1,2].
To reduce harmful exhaust emissions of liquid- Superior mixing properties of this flowfield makes
fuel combustors, without sacrificing their overall it attractive for power generation and aeronautic
performance, an efficient fuel/oxidant mixing pro- propulsion systems, such as low NOx gas turbine,
cess is required. Injection of a liquid jet in a gaseous high speed direct injection (HSDI) diesel engine,
and aircraft engine’s afterburner [3-6]. Many
studies examined the role of different parameters
∗Corresponding author. including liquid properties, operating conditions
E-mail address: madjid.birouk@umanitoba.ca and nozzle geometry in describing spray character-
(M. Birouk). istics of a liquid jet injected into a uniform gaseous

https://doi.org/10.1016/j.proci.2018.08.004
1540-7489 © 2018 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
3238 M. Broumand et al. / Proceedings of the Combustion Institute 37 (2019) 3237–3244

crossflow with negligible velocity fluctuations


(e.g., [1,7-13]). However, there exists no published
research on the effect of a turbulent gaseous cross-
flow’s characteristics, such as turbulence intensity,
on the breakup features of a liquid jet.
Among spray characteristics of a liquid jet in a
gaseous crossflow, the investigation of droplets’ size
and velocity distribution is of a great importance
as they establish the subsequent features of fuel-
oxidant mixing quality, such as droplets’ spatial
distribution, vaporization, mixing rate and conse-
quently the efficiency of liquid fuel combustion. In-
gebo [14] correlated drops mean size, described by
Sauter Mean Diameter (SMD or d32 ), with gas We-
ber number, W eg = ρgu2g d j /σ , gas Reynolds num- Fig. 1. Schematic view of the flow arrangement in the test
ber, Reg = ρg ugd j /μg , and a pressure sensitive di- section.
mensionless group gl/c2 ; where g is the acceleration
due to gravity, l is the mean free molecular path, c
stands for root-mean-square molecular velocity, ρ g stated that, at high-pressure test conditions of up to
and μg are gas density and viscosity, respectively, 7 bar, droplets follow the crossflow velocity within
ug is crossflow velocity, σ is liquid surface tension, a shorter distance compared to that at standard
and dj is the jet diameter. Following this study, Less pressure conditions. Lubarsky et al. [24] studied
and Schetz [15] included liquid jet velocity, vj , in droplets velocity at different Weg , and noted that
their proposed relation for predicting drops SMD. the absolute lag velocity between droplets and
Similarly, Tambe et al. [16] reported that surface the incoming crossflow’s mean velocity is directly
breakup intensifies with increasing jet-to-crossflow proportional to Weg . This observation is confirmed
momentum flux ratio, q = ρl v2j /ρg u2g (where ρ l is by Mashayek et al. [21] who asserted that it takes
liquid density), which results in smaller SMD. a longer distance for droplets to merge into a fast
Zheng and Marshall [17] studied the breakup gas stream.
of a liquid jet at low-Weber-number crossflow, All the above studies investigated the effect of
4 < Weg < 16, and ascertained the influence of a uniform gaseous crossflow with negligible turbu-
Weg on drops size distribution. Becker and Hassa lence on spray characteristics of a liquid jet. Over-
[18] reported the absence of a strong correlation be- all, they revealed that ug (or Weg ) is the predomi-
tween droplets size and Weg at high pressure test nantly controlling parameter on droplets size and
conditions extended from 1.5 to 15 bar. Farvardin velocity distribution. The present study, therefore,
et al. [8] studied the effect of liquid viscosity on aims to investigate the effect of a turbulent gaseous
droplets size downstream of a nozzle using differ- crossflow on spray characteristics of a liquid jet.
ent blends of diesel and biodiesel. They noted that
although viscosity can affect the primary breakup 2. Experimental details
modes of a liquid jet, droplets size distribution
remained almost unchanged after the secondary 2.1. Apparatus
breakup in the far-field region. Taking into account
all of the aforementioned parameters, Song et al. The experimental facility used in this study con-
[19] developed a relation for predicting the mean sisted mainly of an open wind tunnel, two perfo-
size of drops as a function of Weg , q, and liquid/gas rated plates with different solidity ratio, and a liq-
density and viscosity ratios to account for the sur- uid injection system. The wind tunnel is designed to
roundings airflow pressure and liquid properties. generate a uniform subsonic airflow in a transpar-
Droplets velocity downstream of a liquid jet in ent test section made of acrylic, which has a square
a uniform gaseous crossflow is also known to be cross-section of 305 mm × 305 mm and a length of
a strong function of the gas mean velocity ug (or 600 mm (Fig. 1).
Weg ) [20]. Size-velocity profiles generally show that The effect of three different crossflow condi-
(e.g., [21]) the larger the droplets are, the slower tions on the spray characteristics of a liquid jet
they move downstream in the crossflow streamwise is examined. One is without grid, and the other
direction. This is confirmed by Eriksson et al. two cases with a grid (i.e., straight-hole perfo-
[22] who reported that larger droplets exhibited rated plates [25,26]) with a different solidity ratio
slower velocity. Elshamy et al. [23] attributed the (S ≡ solid area/total area) installed at the inlet of
variation in the velocity distribution of droplets in the test section (Fig. 1). The case without grid
a high-Weber-number crossflow, Weg ≈ 100-500, to (WT) acts as a datum such that the breakup is
two main factors; the change of Weg , and the size attributed only to the interaction between a uni-
of the wake region behind the liquid jet. They also form crossflow and a liquid jet. The first perforated
M. Broumand et al. / Proceedings of the Combustion Institute 37 (2019) 3237–3244 3239

plate (PS50) has a solidity ratio of S = 50% with the second mirror was focused. For each test condi-
a nominal mesh spacing M = 16 mm, and the tion, 815 images were collected using an exposure
second (PS67) has a solidity ratio of S=67% and a time of 15 μs which was found suitable to freeze
mesh spacing M=20 mm. Both plates have circular droplets (with no blurring) at each frame with a sat-
straight-holes with a nominal diameter of 12 mm isfactory background brightness level. An in-house
and a thickness of 6 mm. developed MATLAB code was used to estimate
Liquid jet was injected into the crossflow in the the droplets size and velocity distributions. The
test section using a tapered-edged nozzle, which has shadowgraph images were binarized and thresh-
a diameter of dj = 0.5 mm and a fixed length to di- olded such that the liquid column, ligaments with
ameter ratio of L/dj = 4. This nozzle configuration a large aspect ratio (not considered as droplets),
was chosen to prevent the promotion of turbulence small drops with less than 5 pixels or a diameter
and cavitation of liquid flow inside the nozzle. This of 0.0789 mm (as they disappeared in the second
allowed to isolate and study only the effect of cross- frame), overlapping droplets, and the early region
flow velocity fluctuation on spray characteristics of of the column breakup were all filtered out (omit-
a liquid jet. The full characterization of the flow in ted). In doing so, the raw images were subtracted
the nozzle was reported elsewhere [27]. The nozzle from the background image (i.e., without droplets).
exit was centered transversely through the test sec- The resultant grayscale images were then converted
tion ceiling and placed 200 mm downstream of the to binary images using a proper threshold. The
test section inlet to avoid the turbulence develop- droplet diameter is determined as the equivalent of
ing region just downstream of the grid. The nozzle a circular image which has similar projected area.
was inserted 24 mm into the crossflow in order to A bin with 0.0376 mm width was considered to
avoid the effect of wall boundary layer on liquid jet classify droplets size. The centroid of a droplet is
breakup. used to track its motion over successive frames in
a single-frame shadowgraph image. The velocity is
2.2. Diagnostics then determined as the distance between the cen-
troid of the same droplet in the first and second
Particle Image Velocimetry (PIV) was used to frame over the time duration between the two suc-
measure the mean velocity and velocity fluctua- cessive frames. The number of images used (815)
tions of the gaseous crossflow. The PIV setup con- was selected to achieve uncertainties less than 5%
sisted of a Nd:YAG laser with a pulse energy of and 10% (within 95% confidence level) for droplets
135 mJ and a repetition rate of 10 Hz, a double- size and velocity, respectively.
frame FlowSense EO 4 M CCD camera of 20.4 fps
at 2048 × 2048 pixels2 resolution. Olive oil drops 2.3. Test conditions
of about 1 μm were used as seeding particles of
the gaseous crossflow. The duration between pulses Ethyl-alcohol (95% by volume) was used
was adjusted between 40 and 60 μs depending on to study spray characteristics of a liquid jet
the selected crossflow velocity. Three thousand im- at ug =10 m/s (or Weg =2.7) and ug =15 m/s (or
age pairs were acquired using a field of view of Weg =5.9). According to the breakup regime map
160 × 160 mm2 . Dantec Dynamic Studio Software of a liquid jet in a uniform crossflow (e.g., [1,28]),
was used for images post processing and determin- this velocity range of crossflow is associated with,
ing crossflow rms velocities. It was performed using respectively, enhanced capillary and bag breakup
an interrogation area of 32 × 32 with 50% overlap regime conditions. Due to the limitations of the
along with a range validation for eliminating spu- maximum crossflow velocity, we could not reach
rious vectors. In addition, an average filter of 3 × 3 the bag breakup regime with the high solidity ratio
vectors was used for flowfield smoothing. The un- plate (PS67). In addition, due to the geometrical
certainties associated with the PIV measurements constraints of the test section, the maximum
of the mean and fluctuating velocity were within liquid jet velocity was vj, max = 4.5 and 7.6 m/s for,
±1% and ±3%, respectively. respectively, the crossflow velocity of ug =10 and
Shadowgraph technique was used to image the 15 m/s. Ambient temperature and pressure con-
spray and determine its droplets size and velocity ditions were kept standard (298 K and 100 kPa).
distribution. The adopted Z-type setup consisted The properties of ethyl-alcohol are liquid viscosity
of a CMOS high-speed camera with a frame rate of μl = 1.41 × 10-3 Ns/m2 , ρ l = 803.8 Kg/m, and
of 1040 fps and a resolution of 1280 × 1024 pixel2 , σ = 0.0228 N/m. Table 1 lists the test conditions.
a pair of spherical parabolic concave mirrors with
a diameter of 152.4 mm and a focal length of
1143 mm, and a light source. The light beams trav- 3. Results and discussion
eled towards the first spherical mirror, and the re-
flected light beams from the first mirror then trav- 3.1. Crossflow characteristics
eled through the test section towards the second
mirror. The camera was placed on the focal point To investigate the effect of turbulent crossflow
of the second mirror where the reflected light from on spray characteristics of a liquid jet, the gaseous
3240 M. Broumand et al. / Proceedings of the Combustion Institute 37 (2019) 3237–3244

Table 1
Experimental test conditions.
Case ug m/s Weg vj m/s q Grid type
1 10 2.7 1.6 16.5 WT, PS50, PS67
2 10 2.7 3 58 WT, PS50, PS67
3 10 2.7 4.5 130.4 WT, PS50, PS67
4 15 5.9 4.5 60.3 WT, PS50
5 15 5.9 6.3 118.2 WT, PS50
6 15 5.9 7.6 172 WT, PS50

crossflow in the absence of a liquid jet was char- 3.2. Drop size characteristics
acterized using PIV. The spray jets studied here
are considered to be diluted as the ratio of the In order to examine the effect of the turbulent
volumetric flow rate of droplets to that of the air is intensity of a gaseous crossflow on the distribution
small (on the order of 1 × 10−6 ). Also, only a low- of drops size, measurements in the spray region
weber-number liquid jets in the enhanced capillary (FOV domain in Fig. 1) were carried out employing
breakup and bag breakup were studied, and no jet Shadowgraph imaging method. Figure 3 depicts
surface breakup, droplets collisions or coalescence the number fraction of droplets which have the
events were observed in the shadowgraph images. same diameter per percent corresponding to the
A similar approach of study on diluted sprays was test conditions in Table 1. As was expected for a low
recently reported in the literature [29-30]. Several value of Weg and q, drops size exhibits a bimodal
experiments were carried out using two different distribution in the measurements domain, which
perforated plates with solidity ratios of S=50% implies that the liquid jet breaks up in the enhanced
(PS50) and S=67% (PS67), and their results were capillary breakup regime. In the breakup regime,
compared with the case without grid (WT). To the oscillation of the liquid jet causes the forma-
be able to study solely the difference in terms of tion of a neck between the thickened regions (large
velocity fluctuation level, the mean flow conditions droplets) along the liquid column which produces
were kept nearly constant at all test conditions by small size satellite drops after the column breakup.
adjusting the frequency drive of the blower of the For the crossflow with ug =10 m/s (Fig. 3a–c), the
wind tunnel. peak of the number fraction of smaller droplets
Figure 2 illustrates the PIV dataset for the cross- increases from 6% to 8% with increasing vj from
flow streamwise mean velocities, ug =10 and 15 m/s, 1.6 m/s to 4.5 m/s; while the peak of the larger
and the root-mean-square of their corresponding droplets remains almost unchanged. On the other
turbulence velocity fluctuations urms , normalized hand, for the crossflow with ug =15 m/s (Fig. 3d–f),
by ug . These plots are for the region confined within the second peak, associated with the larger size
the FOV domain (Fig. 1), where the horizontal di- droplets, gradually diminishes when increasing vj
rection (x) is along the centerline of the test section from 4.5 m/s to 7.6 m/s, whereas the percentage
and the vertical direction (y) starts from the nozzle of the smaller droplets increases up to 18%. This
injection point. Figure 2a and b confirmed that the single peak distribution of the droplets at high
crossflow mean velocity for different perforated velocity of crossflow and liquid jet flow (Fig. 3f) is
plates are similar and constant throughout the an indication of bag breakup regime. In all these
domain (FOV), and the rms velocities remain ap- cases, it is observed that the crossflow turbulence
proximately unchanged with a slight gradual drop intensity has a insignificant impact on the droplets
along the FOV domain due to the fact that it is still size regardless of the associated breakup regime.
in the decaying part of the crossflow turbulence However, as is illustrated in Fig. 3e, it changes
[25,26]. This implies that the injection of a liquid the basic mechanism of the liquid jet breakup
jet at a 200 mm downstream of the test section’s by delaying the transition of the liquid primary
inlet (Fig. 1) is an appropriate location to have a breakup regime from enhanced capillary to bag
nearly homogeneous turbulence in the horizontal breakup mode. According to Fig. 3e, at ug =15 m/s
direction throughout the FOV domain. Figure 2c and vj = 6.3 m/s, drops size distribution in the uni-
and d shows that ug and urms are also homogeneous form crossflow (WT) shows a single peak profile
in the vertical direction along the liquid injection with a maximum of 16%, while the distribution
point location, x=0. Figure 2b and d shows that, of drops in the turbulent crossflow (PS50) is bi-
using perforated plats with S=50% (PS50) and modal with 8% for the first peak and a smaller
S=67% (PS67), the crossflow turbulence intensity second peak (by about 2%) than the previous
increases from 1.5% (for no grid case) to about cases (Fig. 3a–d).
5.5% and 8.5%, respectively. Comparing ug plots The minimum and maximum number of
(Fig. 2a and c) with those of urms (2b and 2d), it droplets sampled from all 815 images of each case
can be clearly noticed that, contrary to the solidity in Fig. 3 (Table 1) is presented in Table 2. A clear
ratio (S), ug does not have a significant effect on jump in the number of sampled droplets can be
the crossflow turbulence intensity. observed for the cases in bag breakup regime with
M. Broumand et al. / Proceedings of the Combustion Institute 37 (2019) 3237–3244 3241

Fig. 2. Crossflow streamwise mean velocity, ug , and normalized rms urms / ug in x-direction at y =152.5 mm (a) ug ,
(b) urms / ug , and y-direction at x =200 mm (c) ug , (d) urms / ug .

Table 2
The min./max. number of droplets sampled for all the cases in Fig. 3.
Case No: 1 2 3 4 5 6
WT 23/41 22/42 19/29 30/75 239/357 246/358
PS50 27/48 24/45 9/28 31/79 49/141 246/422
PS67 20/46 28/55 19/49 - - -

a single peak distribution (i.e., cases 5-WT and from 10 to 15 m/s, which is still in the enhanced
6-WT and 6-PS50). capillary breakup regime. On the other hand, for
To assess the effect of the crossflow turbulence the case No. 6 (Fig. 3f), which is in the bag breakup
intensity on the mean diameter of a collection of regime, this value decreased to SMD/d j = 1.74. In
different-size droplets under different test condi- all the aforementioned cases, similar to the drops
tions (see Table 1), the SMD of the droplets in the distribution profiles, the turbulent crossflow does
FOV domain depicted in Fig. 1 was also calculated not affect SMD at each specific test conditions
and tabulated in Table 3. This value for the first compared with the uniform crossflow. Nonethe-
three test conditions (case No. 1 to 3 of Table 1) less, in the case No. 5 (Fig. 3e), SMD is similar
was SMD/d j = 2.46 (Figs. 3a–c). This value de- to that of the case No. 6 (SMD/d j = 1.74) for
creased slightly to SMD/d j = 2.20 (case No. 4; the uniform crossflow, while it is the same as that
Fig. 3d), when increasing the crossflow velocity of case No. 4 (SMD/d j = 2.2) for the turbulent
3242 M. Broumand et al. / Proceedings of the Combustion Institute 37 (2019) 3237–3244

Fig. 3. Number fraction of droplets per percent versus drops diameter in the FOV domain for all test conditions reported
in Table 1.

Table 3
SMD absolute values in (mm) for all test conditions re-
ported in Table 1.
Case No: 1 2 3 4 5 6
WT 1.18 1.18 1.26 1.14 0.84 0.9
PS50 1.21 1.21 1.32 1.09 1.08 0.86
PS67 1.2 1.22 1.29 - - -

crossflow. Again, this is an evidence of the role


Fig. 4. Visualization of a liquid jet injected at vj =6.3 m/s
of the turbulence of a crossflow on the breakup
into a (a) uniform and (b) turbulent gaseous crossflow
regime of a liquid jet. with ug =15 m/s.
In order to shed more light on the effect of
the crossflow turbulence intensity on the primary
breakup mechanism of a liquid jet, Shadowgraph
images of case No. 5 (Fig. 3e) for both the uniform (i.e., displaces normal to its axis) in a turbulent
(WT) and a turbulent (PS50) crossflow are depicted more than in a uniform crossflow, with an increase
in Fig. 4. It is observed that the liquid jet injected in the wavelength of the instabilities along the jet’s
into a uniform crossflow (Fig. 4a) breaks up in the column. In the transition region, therefore, the tur-
bag breakup mode [31]. Bags membrane breaks bulent crossflow establishes a condition in which
up into small droplets and the rim of bags disinte- the presence of long wavelength instabilities on
grates into comparatively larger droplets. However, the jet’s surface along with surface tension forces
the turbulent crossflow (Fig. 4b) shapes the column cause the liquid column to break up in the capillary
into an arc structure [32], leading to the breakup mode rather than in the bag breakup, where the
of the liquid jet in an enhanced capillary mode as aerodynamics drag forces of the crossflow are more
ligaments and droplets are generated, which are effective. Nonetheless, increasing further the liquid
larger than drops formed in the bag breakup mode. jet velocity according to case No. 6 (Fig. 3f), it can
As was observed from the Shadowgraph images, be observed that bag like-structures are formed on
the velocity fluctuations of a turbulent crossflow the liquid column in the case of both uniform and
cause the continuous liquid column to fluctuate turbulent crossflow conditions.
M. Broumand et al. / Proceedings of the Combustion Institute 37 (2019) 3237–3244 3243

Fig. 5. Droplet mean velocity in x-direction, uD , and y-direction, vD , versus drops diameter in (a) enhanced capillary and
(b) bag breakup regime.

3.3. Drop velocity characteristics Table 4


Stokes number of droplets with the maximum number
Drops velocity is another important character- fraction for all the cases in Fig. 5.
istic of spray. Hence, the effect of crossflow turbu- Case Grid Type tt (ms) D (mm) tD (ms) St
lence intensity on drops velocity distribution is also
3 WT 0.47 0.29 2.24 4.73
examined. Droplets mean velocity uD in x-direction PS50 0.65 0.24 1.54 2.39
and vD in y-direction are plotted versus their di- PS67 1.23 0.32 2.66 2.17
ameters for both enhanced capillary breakup (case 6 WT 0.29 0.22 1.21 4.15
No. 3) in Fig. 5a, and bag breakup regime (case PS50 0.52 0.22 1.30 2.58
No. 6) in Fig. 5b. As expected, for all test condi-
tions and in both breakup regimes, large droplets
have a smaller ranges of uD than their counter- by further increasing the turbulence intensity to
parts’ small droplets due to their larger Stokes num- 8.5% (in the case PS670), the higher degree of
ber, while they have a higher value of vD owing unsteadiness of the liquid column increases insta-
to their higher y-component momentum. In addi- bilities with longer wavelengths, and hence leading
tion, vD of large droplets is greater than their uD , to a shorter column breakup length. This results in
whereas the opposite velocity trends are observed a reduction in the y-component momentum of the
for small droplets. This implies that, in contrast droplets with a smaller ranges of vD compared to
to small droplets, large droplets do not follow the the two aforementioned cases. In the bag breakup
crossflow easily and consequently penetrate farther regime (Fig. 5b), for which most droplets are less
into the crossflow (which also can be observed in than 0.5 mm, the crossflow associated with PS50
Fig 4). causes a decrease and an increase in uD and vD ,
The effect of crossflow turbulence intensity on respectively. This again indicates that droplets
droplets mean velocity distribution of a liquid jet, generated by the interaction of a liquid jet with a
as is illustrated in Fig. 5, depends on the breakup turbulent crossflow do not necessarily follow the
regime of a liquid jet. In the enhanced capillary crossflow as opposed to a uniform crossflow.
breakup mode (Fig. 5a), the crossflow turbulence Stokes number for droplets with the maximum
intensity associated with the flow cases of PS50 number fraction (Fig. 3) were estimated for each
and PS67 causes a reduction in the ranges of uD condition of Fig. 5. It is defined as St = tD /tt ,
particularly for large size droplets, while they have where tD = (4ρl D )/(3ρg CD (ug − uD ) ) and tt are
the opposite effect on the value of vD . By observing the droplets response time and turbulent time scale
Shadowgraph images, it is observed that increasing of the airflow, respectively, and CD is the drag
turbulence intensity from 1.5% to 5.5% for the PS50 coefficient of a droplet with a diameter of D [29].
case causes the liquid column to fluctuate more Turbulent time scale, tt , is estimated by dividing
in the crossflow, which hinders the liquid column the integral length scale by the mean velocity, ug ,
bending into the crossflow. Such unsteadiness of based on Taylor’s frozen turbulence hypothesis
the liquid column generates large droplets, which since urms is less than 10% of the mean velocity,
does not necessarily follow the turbulent crossflow. ug [25]. As shown in Table 4, for both cases 3 and
This implies a lower uD and a higher vD for the 6, Stokes number of droplets is larger than unity,
droplets in the turbulent crossflow (PS50) com- which implies that droplets are not able to properly
pared to the uniform crossflow (WT). However, respond to turbulent fluctuations.
3244 M. Broumand et al. / Proceedings of the Combustion Institute 37 (2019) 3237–3244

4. Conclusions flow of Air, in NASA Glenn Research Center Report


NASA/CR-2000-210467, 2000.
Droplets size and velocity distribution of a liq- [7] M. Birouk, C.O. Iyogun, N. Popplewell, At. Sprays
uid jet injected into a turbulent crossflow is exper- 17 (3) (2007) 267–287.
[8] E. Farvardin, M. Johnson, H. Alaee, A. Martinez,
imentally investigated. Turbulence intensity of the
A. Dolatabadi, J. Propuls. Power 29 (6) (2013)
crossflow is increased up to about 5.5% and 8.5% 1292–1302.
by changing the solidity ratio of a perforated plate [9] M. Eslamian, A. Amighi, N. Ashgriz, AIAA J. 52 (7)
from 50% to 67%. For the range of test conditions (2014) 1374–1385.
explored in this study, the resulting distribution of [10] M. Broumand, M. Birouk, AIAA J. 54 (5) (2016)
droplets size for each specific breakup regime is 1499–1511.
found independent of the crossflow turbulence in- [11] A.R. Osta, K.A. Sallam, J. Propuls. Power 26 (5)
tensity. However, a turbulent crossflow delays the (2010) 936–946.
transition of the liquid jet breakup from enhanced [12] M. Broumand, G. Rigby, M. Birouk, Flow Turbul.
Combust. 99 (1) (2017) 153–171.
capillary to bag breakup regime. The crossflow
[13] M. Wang, M. Broumand, M. Birouk, At. Sprays 26
turbulence affects droplets velocity distribution of (11) (2016) 1083–1110.
the spray plume by influencing the formation and [14] R.D. Ingebo, J. Propuls. Power 1 (2) (1985) 137–142.
growth of the instabilities on the liquid jet’s surface. [15] D.M. Less, J.A. Schetz, AIAA J. 24 (12) (1986)
Overall, it can be concluded that, while Weg and q 1979–1986.
are known to be the predominantly controlling pa- [16] S. Tambe, S. Jeng, H. Mongia, G. Hsiao, in: 43rd
rameters of the primary breakup regime of a liquid AIAA Aerospace Sciences Meeting and Exhibit -
jet in a crossflow, the role of turbulence intensity of Meeting Papers, 2005.
a crossflow in the transition of the breakup regimes [17] Y. Zheng, A.W. Marshall, At. Sprays 21 (7) (2011)
575–589.
and its impact on spray characteristics should not
[18] J. Becker, C. Hassa, At. Sprays 12 (1–3) (2002) 49–68.
be overlooked. It was not possible to change the dy- [19] J. Song, C. Cary Cain, J. Guen Lee, J. Eng. Gas Tur-
namic range of the crossflow’s turbulence studied in bines Power 137 (4) (2014) 41502.
this experiment, and thus, it would be helpful to use [20] T. Inamura, N. Nagai, J. Propuls. Power 13 (2) (1997)
some high fidelity simulations to complement the 250–256.
present experimental data, as well as extending this [21] A. Mashayek, M. Behzad, N. Ashgriz, AIAA J. 49
study by including higher weber-number breakup (11) (2011) 2407–2420.
regimes such as shear breakup. [22] P. Eriksson, R. Orbay, J. Klingmann, in:
ICLASS-2006, 2006.
[23] O. Elshamy, S. Tambe, J. Cai, S.-M. Jeng, in: 45th
AIAA Aerospace Sciences Meeting and Exhibit,
Acknowledgments American Institute of Aeronautics and Astronautics,
2007.
This research was funded by the Natural [24] E. Lubarsky, J.R. Reichel, B.T. Zinn, R. McAmis,
Sciences and Engineering Research Council of J. Eng. Gas Turbines Power 132 (2) (2009)
Canada (NSERC). 21501.
[25] R. Liu, D.S.-K. Ting, G.W. Rankin, Exp. Therm.
Fluid Sci 28 (4) (2004) 307–316.
References [26] R. Liu, D.S.-K. Ting, J. Fluids Eng. 129 (9) (2007)
1164–1171.
[1] M. Broumand, M. Birouk, Prog. Energy Combust. [27] M. Broumand, M. Birouk, Phys. Fluids 29 (2017)
Sci. 57 (2016) 1–29. 113303.
[2] A. Mashayek, N. Ashgriz, Handbook of Atomiza- [28] K.A. Sallam, C. Aalburg, G.M. Faeth, AIAA J. 42
tion, Springer ScienceþBusiness Media, LLC, 2011, (12) (2004) 2529–2540.
pp. 657–683. [29] A. Kourmatzis, A.R. Masri, Exp Fluids 55 (2014)
[3] M. Herrmann, Proc. Combust. Inst. 33 (2) (2011) 1659.
2079–2088. [30] A. Kourmatzis, A.R. Masri, J. Fluid Mech. 764
[4] R. Ragucci, A. Bellofiore, A. Cavaliere, Proc. Com- (2015) 95–132.
bust. Inst. 31 (2) (2007) 2231–2238. [31] C.-L. Ng, R. Sankarakrishnan, K.A. Sallam, Int. J.
[5] J.M. Desantes, J. Arrègle, J.J. López, J.M. García, Multiph. Flow 34 (3) (2008) 241–259.
Fuel 85 (14–15) (2006) 2120–2132. [32] G. Vich, M. Ledoux, Int. J. Fluid Mech. Res. 24 (1–3)
[6] M. Leong, V. McDonell, G. Samuelsen, Mixing of an (1997) 1–12.
Airblast-Atomized Fuel Spray Injected into a Cross-

You might also like