ArticlePDF Available

Spatiotemporal Characterization of Wave-Augmented Varicose Explosions

Authors:
1
Spatiotemporal Characterization of Wave-Augmented Varicose Explosions
1
2
D. M. Wilson1, W. Strasser1, R. Prichard1
3
4
1School of Engineering, Liberty University, Lynchburg, VA 24515, USA
5
6
ABSTRACT
7
The onset of three-dimensional instabilities during -(WAVE) liquid
8
disintegration is characterized for an annular flow of a shear-thinning slurry that is fed into a central transonic
9
steam flow. Droplet production inside the nozzle is enhanced by ligaments radially flicking up from the slurry
10
wave into the steam flow with radial:axial velocity ratios exceeding 0.5. The wave also leaves residual
11
ligaments in its wake, which facilitate further disintegration. After birth, a wave spends 80% of the wave cycle
12
period building up to peak height at the nozzle exit. Two effervescent mechanisms are provided as 1) steam
13
penetrates the rising wave and surface deformation allows steam fingers to force through, and 2) the wave
14
collapses on itself, trapping steam. Baroclinic torque drives the development of Rayleigh-Taylor (RT)
15
instabilities and reaches values on the order of  1/s2. Both RT and Kelvin-Helmholtz instabilities are
16
self-amplified in a viscosity-shear-temperature instability cycle because the slurry is non-Newtonian.
17
Velocities inside the nozzle (wave formation region) are generally azimuthally similar (two-dimensional), but
18
those outside the nozzle (radial bursting region) are azimuthally uncorrelated (three-dimensional). Inter-
19
variable correlations show significant decoupling of quantities beyond the nozzle exit, and local strain rate
20
fluctuations were found to correlate particularly well with bulk system pulsation. Although adaptive mesh
21
refinement (AMR) can provide computationally efficient resolution of gas-liquid interfaces, this technique
22
produced different results than an equivalent non-dynamic mesh when modeling WAVE. Gradients were
23
particularly affected by AMR, and turbulent kinetic energy showed differences greater than 150% outside the
24
nozzle.
25
Keywords: CFD, multiphase, transonic, instability, adaptive mesh refinement
26
The author to whom correspondence may be addressed: wstrasser@liberty.edu@liberty.edu
27
2
1. INTRODUCTION
28
The relevance of atomization for modern life cannot be understated. Those atomization processes found in nature
29
designs have been successfully exploited for applications in many industries, such as automotive, aerospace,
30
chemical, and agricultural. Though widely employed, atomization is a complex and diverse process that continues
31
to be studied with great vigor. Various fluid instabilities lead to atomization, and the nature of droplet formation
32
makes atomization an inherently three-dimensional (3D) process. The Kelvin-Helmholtz instability (KHI),
33
induced by velocity gradients, is often a contributor to atomization but is only two-dimensional (2D). For a moving
34
surface, KHI operates in the longitudinal direction to excite the interface, creating perforations, tongues, and
35
waves. Variation in the transverse direction can arise from Rayleigh-Taylor instabilities (RTI), which are induced
36
by a misalignment between the density and pressure gradient vectors; RTI arise from the extra term



37
in the vorticity equation, where is density, is pressure, and is the direction vector. Additionally, surface
38
curvature can induce Rayleigh-Plateau capillary instabilities (RPI). KHI, RTI, and RPI can be important and
39
simultaneous players in the disintegration of liquids.
40
41
Controlled atomization processes utilize a range of methods. Some common types of atomization-enhancing
42
techniques include swirl,1-3 effervescence,4, 5 and an assisting gas stream.6-8 Twin-fluid atomization utilizes a high-
43
speed co-lowing gas stream to disrupt the liquid. Typical designs involve a central liquid stream surrounded by a
44
co-axial gas flow. The gas-liquid shear encourages destabilization of the bulk liquid for disintegration and droplet
45
production. Gas-Centered Swirl Coaxial (GCSC) injector designs have been used for propellant atomization with
46
inverted the feeds.9 A central gas flow is then surrounded by a co-axial liquid flow, and pulsing characteristics can
47
develop at certain gas-to-liquid momentum ratios.10 Alekseenko et al. studied disturbance waves in vertical
48
annular liquid flow around a central gas stream, but these waves are small relative to the nozzle.11
49
50
A Newtonian fluid is typical for atomization processes, and non-Newtonian fluids will alter atomization
51
characteristics. Although KHI, RTI, and RPI are, fundamentally, inviscid instabilities, variations in viscosity can
52
both 1) modify the growth of these perturbations and 2) further destabilize the interface. Non-Newtonian fluid
53
layers can be prone to instability even when operating at low Reynolds numbers.12 When studying a shear-thinning
54
3
fluid on a slope, Millet et al. reported increased celerity and likelihood of instability compared to the Newtonian
55
counterpart.13 A unique breakup mode for secondary atomization was discovered in the study of non-Newtonian
56
coal water slurries.14 Guo et al. report that the central gas flow promotes instability of high-speed annular power-
57
law fuel jets and that a thinner fuel film and higher gas density also encourage breakup.15 Both inverted feeds and
58
an ing viscous, non-
59
Newtonian waste slurries.16 Applications could include waste-to-energy conver   
60
technology17 and gelled propellant atomization.18-21
61
62
It has been shown that a certain inverted-feed, forced interaction atomizer design produces periodic, high-
63
blockage-ratio waves inside the nozzle when a viscous, shear-thinning fluid present.22 Two past numerical
64
studies have investigated aspects of this phenomenon with hot, subsonic steam as the assisting gas and banana
65
puree as the working fluid.22, 23 Banana puree is both viscous and shear-thinning, with well-recorded viscosity
66
data in the literature.24 WAVE atomization could be instrumental for manure slurries (energy reclamation) and
67
gelled propellants. A predominantly 2D analysis revealed basic mechanisms driving liquid wave formation
68
and collapse.22 An investigation of atomization downstream of the nozzle highlighted the importance of waves
69
for transonic disintegration of puree via periodic radial bursting.23 Since the characteristic wave cycling leads
70
to bursting (that in turn enhances atomization), the process has been termed -Augmented Varicose
71
  However, major gaps in understanding the WAVE phenomenon remain. Most
72
importantly, the 3D nature of the interior waves and related instability cycle are almost entirely unexplored.
73
Among the many questions that arise are the following: Where and how do RTI form in the wave prior to
74
rupture, and is enough time in a wave cycle allotted for RTI to develop? To what degree are waves azimuthally
75
uniform? If not, where does the uniformity break down? What mechanism drives droplet breakaway inside the
76
nozzle before the radial burst is it wave stripping or wave flicking? Do quantities at all locations fluctuate
77
with wave cycles? Do flow metrics correlate with one another at various times and locations?
78
79
Adaptive mesh refinement (AMR) was considered as a numerical technique, but initial results showed AMR
80
to be an outlier to numerical trends.22 Because of the complexity and small scales in these atomization
81
4
phenomena, it can be computationally expensive to resolve the explicitly gas-liquid interface using CFD. To
82
increase computational efficiency, AMR has been used been used to model both waves and atomization.25, 26
83
Rather than refine all cells within the computational domain, AMR locally and dynamically refines the mesh only
84
around the interface as the simulation proceeds. Despite the benefits of AMR, the aforementioned results indicate
85
a need to further understand how AMR affects simulations. It remains unclear how AMR differs qualitatively and
86
which quantities are most affected by AMR.
87
88
In this paper, a numerical study is presented to reveal for the first time the spatiotemporal aspects of WAVE;
89
azimuthal instabilities are a primary focus. Careful evaluation of adaptive mesh refinement (AMR) and how it
90
affects the modeling of banana puree slurry atomization is also an important contribution. Our new findings are
91
communicated in four distinct sections, each of which seeks to address unanswered questions from previous
92
studies. Section 3.1 includes cartesian (unraveled from cylindrical) pictures and animations of the wave (so that
93
the azimuthal variation and onset of RTI is evident), as well as calculations of baroclinic torque and discussion of
94
the liquid viscosity-shear-temperature instability cycle. Section 3.2 presents contours of radial versus axial
95
velocity to elucidate mechanisms for droplet breakaway from the liquid wave inside the nozzle. Section 3.3
96
provides extensive frequency and correlation analyses across 100 spatially diverse signals, including
97
demonstration of Fast Fourier Transform (FFT) convergence. An analysis of wave cycle timing is also included,
98
showing leading and lagging responses. Finally, Section 3.4 presents a qualitative and quantitative assessment to
99
clarify the effect of AMR on WAVE simulations.
100
101
2. METHODS
102
2.1 Computational Methods
103
The governing Navier-Stokes equations, formulated for multiphase flow in vector notation, are presented in
104
Equations 1-3. Symbols in Equations 1-3 are defined as follows: is time, is the velocity vector, is the static
105
temperature, is phase volume fraction, is density, is constant pressure heat capacity, is laminar
106
conductivity, is the turbulent viscosity,  is the turbulent Prantdl number, is pressure, is gravity, is the
107
surface tension force vector, is the laminar shear stress tensor, and is turbulent shear stress tensor. Properties
108
5
are arithmetically phase-averaged, the banana puree slurry is modeled as incompressible, and the ideal gas equation
109
of state is used to compute steam density.
110
111
󰇛󰇜󰇛󰇜 (1)
112
113
󰇛󰇜  󰇛󰇜
(2)
114
115
󰇛󰇜󰇟󰇛󰇜󰇠 󰇣󰇡
󰇢󰇤 (3)
116
117
The compressible Reynolds-Averaged Navier-Stokes and volume-of-fluid (VOF) equations were discretized and
118
solved using double precision segregated ANSYS Fluent 2020R1 software. The gas-liquid interface was
119
reconstructed explicitly by means of the geometric reconstruction technique (also known as piecewise linear
120
interface capturing or PLIC).27, 28 Turbulent effects were included via a homogeneous shear stress transport (SST)
121
k-model. SIMPLE (Semi-IMplicit Pressure Linked Equations) was used for pressure-velocity coupling, and
122
mostly second order discretization stencils were employed. With the segregated approach of SIMPLE, pressure
123
and velocity are updated sequentially instead of simultaneously.29 Time step was varied to preserve a Courant
124
number of 1 throughout the entirety of the simulation. For the finest mesh, the time step was generally on the
125
order of one-hundred-thousandth of a wave/pulsation cycle (1×10-8 s).
126
127
The Herschel-Bulkley model is used to describe the shear-thinning (beyond a yield stress) and temperature-
128
thinning nature of the slurry. Banana puree viscosity is modeled as a function of both strain rate and temperature
129
according to the data of Ditchfield et al.24 A user define function (UDF) in Fluent computes viscosity () according
130
to Equations 4-7 from strain rate magnitude (󰇗) and temperature ( in °C) values. In Equation 4, is the yield
131
stress, and are calculated according to Equations 6 and 7,
󰇗  is the lower strain rate bound, and
132
is the corresponding upper viscosity bound. Although our implementation method used herein matches that
133
of the validated method used in Ref. 30,30 we sought to further verify our calculations. Instantaneous cell-
134
6
centered values of strain rate and temperature were collected at five locations (to be discussed more in a future
135
section) within the model for a single moment in time. The viscosity observed in Fluent matched excellently
136
that calculated by hand with Equations 4-7 (less than 0.005% difference), verifying correct implementation of
137
the viscosity UDF.
138
139
󰇫 󰇗
󰇗
󰇗󰇗 󰇗 󰇗 (4)
140
141
 (5)
142
143
 (6)
144
145
 (7)
146
147
Extensive validation of the methods here employed for transonic wave formation and atomization has already
148
been conducted over the course of the last 10 years.27, 31-38 The SST k-turbulence model, which is employed
149
here, was sufficient to reveal important physical mechanisms. Experimental results were reproduced both
150
quantitatively and qualitatively, and the primary validation exercise are summarized as follows. First,
151
computations revealed the globally pulsing nature of an industrial three-stream air-water atomizer as
152
qualitatively observed in experiments. Second, the experimental acoustic signature of pulsations and primary
153
atomization ligament wave positions were quantitatively reproduced with numerical simulations. Third, the
154
axial droplet size distribution in a non-Newtonian injector from experiments aligned quantitatively with
155
numerical results. Additionally, assessment of droplet size distribution did not significantly alter with changing
156
azimuthal angle, even for 1/32nd of a full 360° azimuth. Fourth, the numerical trajectory of a disintegrating
157
droplet (after exposure to a normal shock wave) matched the analytical trajectory for said droplet. Furthermore,
158
broadly similar atomization systems have been studied (recessed with a high gas-to-liquid momentum ratio
159
and an inverted feed), and we find that the results presented here correspond to the globally varicose (but
160
7
locally sinuous from the Lagrangian perspective of the flapping annular waves) pulsing nature expected of
161
such systems.10
162
163
Models were considered to be at quasi steady state (QSS) when various point monitor signals (discussed later)
164
were statistically stationary. Unless otherwise noted, all results presented in this paper are using QSS data, and
165
quantities were only time-averaged across QSS data. Since the system experiences bulk pulsation, flow time is
166
normalized by the time elapsed between pulses. This is referred to as or, equivalently, as a
167
. Waves and the general pulsing phenomenon are operating together in the same cyclic pattern.
168
The model used for the majority of results, Ref-3, was run for 12 PTs of QSS data and includes 8.5 PTs of time-
169
averaged data. A recently developed methodology was used to optimize the hardware utilization.39
170
171
2.2 Mesh and Boundary Conditions
172
A summary of the geometry, mesh, and boundary conditions is provided here. Rather than the typical twin-fluid
173
atomizer design with a central liquid flow surrounded by a coaxial gas flow, the streams are reversed. An outer
174
slurry annulus surrounds a central steam flow. The slurry pool is exposed to the hot subsonic center steam flow
175
before the nozzle exit, and the nozzle orifice is extended to encourage significant steam-slurry interaction before
176
exiting the nozzle. The full 360° domain is simulated by a 90° azimuthal slice bookended by periodic boundary
177
conditions. Changing the angle from 45° to 90° produced no noticeable change in the axial droplet size profile,23
178
and the sufficiency of the 90° azimuth will be further discussed among the results presented here. The downstream
179
atomization domain spans 1.5 nozzle diameters in the radial direction and extends 2 nozzle diameters in the axial
180
direction. Constant mass flow rates of 0.021 kg/s for steam and 0.79 kg/s for the slurry resulted in a gas-liquid
181
mass ratio of 2.7%. Inlet temperatures for the slurry and steam were set to 304 K and 393 K, respectively. The
182
steam inlet feed turbulence is set by defining a turbulent kinetic energy () and specific dissipation rate (), where
183
= 48 m2/s2 and = 2.5×105 1/s. Both and are spatially constant and were chosen arbitrarily. All results
184
presented in this paper should be understood with these inlet conditions in mind, and other inlet conditions could
185
be studied to determine the influence of GLR and turbulence feed conditions on wave formation and atomization.
186
187
8
Figure 1 provides an overview of the geometry and shows an example of the Ref-3-AMR mesh. 
188
represent the radial, axial, and azimuthal directions, respectively. In total, six distinct meshes were evaluated. A
189
Base mesh was refined n times to produce Ref-1, Ref-2, and Ref-3 meshes. The total element count increases by
190
a factor of nearly eight during refinement, as each cell length is cut in half. Note that the Base mesh was not refined
191
farther back in the slurry annulus and steam pipe, but all regions of gas-liquid interaction were included in the
192
refinements. The additional two meshes are Ref-2-45, which replaces the 90° azimuth with a 45° azimuth, and
193
Ref-3-AMR, which utilizes adaptive mesh refinement (AMR). AMR provides the same refinement level as Ref-3
194
but increases computational efficiency by refining dynamically only around the gas-liquid interface while the rest
195
of the domain remains at a Ref-2 refinement level. To closely track the gas-liquid interface, the mesh is refined
196
every 5 time steps, and a stringent VOF gradient criterion is used. Ref-3-AMR maintained around 28 million
197
elements compared to the 132 million elements in the Ref-3 mesh.
198
199
The Ref-3 mesh was found to be reasonably mesh independent across a range of metrics, including turbulent
200
kinetic energy, velocity, and slurry volume fraction at the nozzle exit.22 Furthermore, wave physics comparable to
201
those produced by Ref-3 (pulsation frequency and wavelength, for example) were revealed by Ref-1, two
202
refinement levels below Ref-3. We acknowledge that all relevant length scales are not resolved, but Ref-3
203
maintains the requirement of at least four cells across each droplet,23 which was shown to be sufficient for viscous
204
slurry atomization using our computational methods.40 Strasser and Battaglia provide more discussion on
205
implications of mesh resolution for atomization.31 The trend of numerical results versus mesh size showed Ref-3-
206
AMR to be an outlier. For these reasons, Ref-3 is used for all results presented in this paper unless otherwise noted.
207
The difference in numerical output from the Ref-3 and Ref-3-AMR meshes is an important point; a comparison
208
of the two meshes will be presented later. Figure 2 presents a side view of the pre-filming (wave formation) region
209
of the Ref-3 mesh. The region where slurry (this is where slurry
210
waves rise into the steam flow) and the extension of the nozzle orifice 
211
before exiting the nozzle.
212
213
9
214
215
Figure 1 Oblique view of a representative Ref-3-AMR surface mesh, illustrating the nature of adaptive mesh
216
refinement (AMR). Starting with a Ref-2 mesh, the cells around the gas-liquid interface are dynamically
217
refined every 5 time steps. Through a given pulsing cycle, Ref-3-AMR fluctuates between ~26 and ~30 million
218
elements compared to the 132 million Ref-3 mesh. Both Ref-3 and Ref-3-AMR use a 90° azimuth with periodic
219
boundary conditions on either side. The radial, axial, and 
220
respectively.
221
222
223
224
225
Figure 2 Side view of the 132 million element Ref-3 mesh for the pre-filming (wave formation) region of the
226
atomizer. A central steam flow is interrupted by the injection of the slurry, forcing the two phases to interact
227
before exiting the nozzle. slurry and steam meet) and travel across
228

229
230
231
10
3. RESULTS AND DISCUSSION
232
3.1 3D Wave Cycle
233
The inverted-feed, forced interaction nozzle design leads to highly regular bulk pulsations in the system. These
234
pulsations may be looked at from two perspectives: inside and outside the nozzle. We will examine both in
235
this paper. Inside the nozzle, annular waves form at regular intervals (corresponding to bulk system pulsation),
236
rising out of the wave pool (labeled in Figure 2). One wave forms 
237
toward the center of the steam flow and collapses while exiting the nozzle. Previous studies have mostly
238
considered 2D aspects of these waves (KHI, which is integral to the wave formation process, is 2D),22 leaving
239
important 3D characteristics uninvestigated (RTI, for example, manifests in the azimuthal dimension). We
240
seek to clarify how azimuthal variation develops in the wave and leads to atomization downstream.
241
242
Outside the nozzle, regular radial bursting of slurry is observed. This bursting manifests as a three-part
243
sequence: 1) stretch, 2) bulge, and 3) burst. While a wave is forming inside the nozzle, an annular sheet of
244
slurry stretches from the nozzle exit. As the wave collapses, the windward pressure build-up propagates
245
through the wave, and the slurry sheet bulges radially. Finally, the slurry sheet ruptures altogether. Both wave
246
cycling and radial bursting occur at a frequency around 1000 Hz (more on this later). The wave formation is a
247
sinuous instability manifestation (from a Lagrangian perspective of an observer moving with the wave leaving
248
the nozzle), but the bulk pulsation outside of the nozzle (radial bursting) is a varicose manifestation. Wave
249
formation and collapse are integral -by forcing the wave to crash
250
into itself and enhance the disintegration of viscous, non-Newtonian fluids with radial explosions. The Weber
251
(), Reynolds ( ), and Strouhal ( ) numbers characterizing the wave are
252
2.0×104, 1.2×104, and 0.08, respectively. Here, is wavelength,  is wave speed, is wave frequency, is the
253
steam-slurry bulk velocity, is slurry density, is surface tension, and is the lower end of slurry viscosity.
254
255
To better understand cross-wave variation during the 3D wave cycle, Figures 3-6 present a unique view of the
256
annular wave: the 90° azimuthal slice has been unraveled to    (rather than annular)
257
visualization of the slurry interfacial motion with a superimposed contour in the background. The result is
258
11
more comparable to what ocean waves on a beach look like and provides insight into 3D aspects of the wave
259
physics. Wave formation is illustrated in frames 1-3 and wave collapse in frames 4-6. Figure 3 (Multimedia
260
view) shows the unraveled slurry surface colored by viscosity, and the background contour is colored by strain
261
rate. This view is from the perspective of inside the nozzle to observe wave formation. High strain-rate regions
262
appear as the hot steam contacts the slurry both at the wave and with droplets in the free stream. Viscosity
263
reduces according to the shear-thinning nature of the slurry where strain rate is highest and the temperature-
264
thinning effect by the hotter steam. Viscosity reduction causes ligaments to flick up into the steam,
265
compounding the shear-thinning effect. Some azimuthal variation in viscosity is evident, although spatial
266
variability fluctuates throughout the wave cycle. Azimuthal variation in the slurry is present, though minimal,
267
in frames 1-3 but becomes very pronounced as the wave crashes in frames 4-6. Frames 4 and 5 show irregular
268
valleys and ridges forming across the azimuth. The RTI time scale, as approximated with a low-Re method41
269
for the low end of the slurry viscosity spectrum, is an order of magnitude lower than the wave time scale. This
270
suggests that sufficient time is available for RTI development.
271
272
273
274
Figure 3 Sequential views of unraveled wave with the slurry surface colored by viscosity and contours of
275
strain rate in the background. Pictures are at equally spaced flow time intervals through one representative
276
wave cycle. The view is from the inside of the nozzle looking down on the wave. The shear- and temperature-
277
thinning nature of the slurry is highlighted by viscosity reduction in response to high strain rate as the wave
278
12
penetrates into the hot steam. Azimuthal variation is present as the wave forms (frames 1-3) but increases
279
significantly as the wave collapses (frames 4-6). (Multimedia view).
280
281
Figure 4 (Multimedia view) presents a similar visualization as that in Figure 3 but viewed from outside the
282
atomizer (looking underneath the wave). This perspective shows the annular slurry sheet (now flattened)
283
stretching out from the nozzle as a wave rises. In this view, the wave is rising underneath the slurry sheet. The
284
point of maximal extension before rupture (frame 4) corresponds to the highest viscosity of the slurry sheet,
285
as it experiences minimal shear outside the nozzle radius. The sheet is destabilized by both azimuthal variation
286
from RTI (heavier slurry being accelerated by the lighter steam) and azimuthal variation in viscosity. Frame 4
287
reveals significant and azimuthally variant bulging underneath the wave, which corresponds to valleys in the
288
wave surface. Clearly, the wave is not impacting the slurry sheet uniformly. We also note that the multiple
289
bulges across this 90° azimuthal slice indicates that 90° is a sufficient angle to capture variation in the azimuth.
290
Parts of the wave contact local portions of the slurry sheet before others-
291
slurry sheet. Frame 5 marks the pre-rupture of the slurry sheet, where small portions begin to break apart. The
292
collapse of the wave and the instability of the slurry sheet both likely contribute to pre-rupturing. The further
293
destabilized slurry sheet completely ruptures in frame 6 in a violent radial burst.
294
295
296
297
13
Figure 4 Sequential views of unraveled wave with the slurry surface colored by viscosity and contours of
298
strain rate in the background. Pictures are at equally spaced flow time intervals through one representative
299
pulsing sequence. The view is from the outside of the nozzle (looking up from underneath the wave) to
300
illustrate the radial bursting phenomenon. As the slurry stretches out in an annular (flattened for these views)
301
sheet, RTI - as the collapsing
302
wave starts to contact the slurry sheet, followed by complete rupture in frame 6. (Multimedia view).
303
304
Waves have a high blockage ratio that significantly affects the steam pressure and velocity via local
305
acceleration; both fluctuate periodically with the pulsing cycles. Figure 5 (Multimedia view) illustrates these
306
effects by showing the unraveled wave surface colored by pressure with a Mach number contour in the
307
background. As the wave rises, it shelters its leeward side from the oncoming steam flow, thereby reducing
308
the flow area for the steam exiting the nozzle. Steam is then accelerated to transonic velocities. Though not
309
explicitly shown in Figure 5, small regions are supersonic. Consequently, steam compresses on the windward
310
side of the wave to build up pressure, and the steam accelerates above the wave crest through the reduced-area
311
opening. Both phenomena have implications for wave formation. The windward high-pressure zone exploits
312
irregularities in the slurry surface caused by RTI and viscosity gradients. The transition from wave formation
313
to wave collapse (frame 3 to frame 4) is significant: pressure increases sufficiently to overcome inertia and
314
surface tension, and the increase in azimuthal variability is marked. We observe in frame 5 the propagation of
315
the windward high-pressure zone axially as the wave collapses, which is a driving force in the radial bursting.
316
317
318
319
14
Figure 5 Sequential views of unraveled wave with the slurry surface colored by pressure with contours of
320
Mach number in the background. Pictures are at equally spaced flow time intervals through one representative
321
wave cycle. The view is from the inside of the nozzle looking down on the wave. The high blockage ratio of
322
the wave leads to pressure build up on its windward side and reduces the exit area for the steam to accelerate
323
it above the wave crest. Steam clearly reaches transonic speeds. Pressure buildup drives wave collapse and
324
exploits azimuthal variations in the slurry surface (caused by RTI and viscosity gradients) for disintegration.
325
(Multimedia view).
326
327
Figure 6 (Multimedia view) provides a side view to of the wave cycle (gray interface) with a contour of
328
temperature in the background. Much like pressure, temperature cycles as the steam periodically compresses
329
on the windward side of the wave. Highest steam temperatures occur in frames 4 and 5 and will contribute to
330
surface destabilization and wave disintegration by reducing slurry viscosity, although shear is undoubtedly the
331
dominant driving force behind viscosity changes. The wave appears to be largely 2D as it rises and approaches
332
the nozzle exit. As it reaches the nozzle exit in frame 4, the wave is transitioning to a more 3D surface. This
333
corresponds to the dramatic increase in azimuthal irregularity as pressure exacerbates existing surface
334
variations and temperature thins the slurry.
335
336
337
338
Figure 6 Sequential views of unraveled wave (gray) with contours of temperature in the background. Pictures
339
are at equally spaced flow time intervals through one representative wave cycle. The view is directly from the
340
side of the wave, effectively outing the wave profile. Wave blockage causes the steam to compress, cycling
341
the temperature with wave formation and collapse. The wave is largely 2D until around the nozzle exit, where
342
it collapses, and azimuthal variations are exacerbated for a strongly 3D surface. (Multimedia view).
343
344
15
It has already been noted that RTI is driven by the baroclinic torque term 󰇡


󰇢 in the vorticity equation.
345
Density-based torque on the interface creates vorticity that will tend to increase the misalignment of pressure and
346
density gradient vectors. This in turn creates additional vorticity, leading to further misalignment. Figure 7
347
(Multimedia view) presents contours of baroclinic torque. The highest values are on the order of  1/s2,
348
indicating significant misalignment of gradient vectors and strong RTI activity in the wave. Baroclinic torque is
349
greatest at the steam-slurry interface, where density gradients are highest. Aside from directly at the interface,
350
baroclinic torque is highest in the regions where droplets are being stripped off the wave and interacting with the
351
steam. This indicates that as the shear layer forms, RTI becomes important very early in the wave life.
352
353
354
355
Figure 7 Sequential side contours of baroclinic torque through one representative wave cycle. Rayleigh-Taylor
356
instabilities (RTI) arise from this baroclinic torque term in the vorticity equation. Baroclinic torque is highest
357
at the interface, where the density gradient is highest. The largest values shown are on the order of 
358
1/s2, indicating significant RTI activity present in the wave from its birth. (Multimedia view).
359
360
RTI and KHI are present early in the wave and lead to minute interfacial deformation. As the approaching
361
steam navigates the newly roughened surface, spatial variations in temperature and strain rate occur across the
362
wave. The shear- and temperature-dependent slurry responds accordingly, producing local variations in
363
viscosity across the surface. Because of viscosity variation, the steam-slurry interface develops axial and
364
azimuthal wavelength spatial variability. Also, due to viscosity variation, the interfacial stress develops spatial
365
variation. Finally, the process is repeated as the surface deformation excites RTI and KHI. Figure 8 illustrates
366
16
the cycle, where T is temperature, SR is strain rate, and WL is wavelength. Note that we are not utilizing linear
367
instability analysis but rather the local momentum balance to reveal these instabilities.
368
369
370
Figure 8 Instability cycle, where the Kelvin-Helmholtz instability (KHI) and Rayleigh-Taylor instability (RTI)
371
cause surface variations that in turn excite the instabilities. Here, T is temperature, SR is strain rate, and WL
372
is wavelength.
373
374
375
3.2 Droplet Production
376
In a previous study, the effect of the wave on droplet production was largely addressed from the perspective
377
of bursting outside the nozzle,23 but questions remain regarding the role of 3D surface instabilities in droplet
378
production inside the nozzle. It has been noted that small droplets break away from the wave inside the nozzle,
379
and this was primarily attributed to steam shear (and later the collapse of the wave).23 A more thorough
380
investigation of this phenomenon is presented here, but we will first put the atomizer and its droplet production
381
mechanisms in context.
382
383
The Wave-Augmented Varicose Explosions (WAVE) design is essentially a combination of a Gas-Centered
384
Swirling Coaxial (GCSC) and Effervescent atomizer which capitalizes on the advantages of both and
385
incorporates an additional element of radial momentum generation. Effervescence is introduced when the wave
386
crashes onto the annular slurry sheet at regular intervals. Steam is sandwiched between the wave and sheet,
387
introducing bubbles into the liquid phase. The Gas-Centered feature of GCSC is preserved, though swirl is not
388
included in the current design. The lack of swirl necessity is a benefit of the current WAVE design because
389
swirl increases internal geometric complexity and could require maintenance of fouled swirl elements. Besides,
390
a previous study showed the impact of swirl in a multi-stream non-Newtonian atomizer to be minimal.33 Future
391
research may include revisiting this idea.
392
17
393
Tat least
394
three mechanisms. 1) The momentum of the crashing wave creates radial bulges in the liquid film, bursting
395
droplets outward at regular intervals. 2) The high blockage ratio of the wave causes an intense pressure increase
396
behind the wave that assists the wave momentum in driving liquid film rupture and slurry disintegration. 3)
397
The extension of the slurry as a wave into the steam flow increases the interfacial area, providing more space
398
for the steam to peel off droplets. 42 refers to the fact that, from a fixed external observatory frame,
399
  
400
refers to the radial blasting of droplets caused by wave crashing.
401
402
The first two mechanisms have been thoroughly investigated.23 It is the third mechanism that will be explored
403
here. Droplets clearly break away from the wave as it penetrates into the steam flow inside the nozzle (before
404
the wave completely collapses), but what factors contribute to this outcome? It remains unclear whether the
405
steam is merely stripping droplets off an axially moving wave surface or if slurry ligaments are being flicked
406
up into the steam with significant radial velocity. To address this question, we present contours of radial
407
velocity divided by the absolute value of axial velocity in Figure 9 and Figure 10 (Multimedia view). The
408
outline of the steam-slurry interface is marked in black. Blue represents movement inward towards the central
409
steam flow, and red represents movement outward toward the beach and beyond. For the purposes of this
410
discussion (and from the perspective of Figure 9 and Figure 10), we will refer to blue regions as moving
411

412
413
The time sequence in Figure 9 reveals the temporal development of radial versus axial velocity through a given
414
wave cycle. By frame 3, the wave is penetrating significantly into the steam flow. Consequently, the wave
415
develops dark blue on its back, corresponding to the deflection of steam upward. This deflection
416
is significant close to the wave but less so further upstream. Just behind the wave jet, much of the slurry is
417
directed axially (white), but the wave jet itself has significant radial thrust pushing upward into the steam flow.
418
Radial velocity is at least 50% of the axial velocity in dark blue regions, which is the limit of the scale. By
419
18
frame 4, where the wave height peaks, much of the wave base is moving in the axial direction. However, we
420
still observe strong upward movement at the top of the wave, and some ligaments are present. In other words,
421
we observe the slinging of ligaments upward as the main body of the wave is moving forward. Steam around
422
the wave tip and ligaments also has strong upward motion. Non-axial flow in an atomizer has been found to
423
correlate with higher droplet production efficiency.31 As an aside, the beginning of the radial bulge is quite
424
evident as the red patch just at the end of the beach in frame 4. The flicking up of ligaments in frame 4 helps
425
explain what we term  . For example, there is a nearly vertical ligament around the
426
nozzle exit in frame 5, which is unexpected because of the high-velocity steam directed toward it. This appears
427
to be a residual ligament derived from the windward ligament in frame 4. We call the windward ligament a
428
primary wave tip. By frame 5, the ligament is moving downward, but
429
it is an enduring form of the upward moving ligament in frame 4, and its position enables steam to disintegrate
430
the slurry more effectively. Curiously, the steam on the leeward side of the residual ligament is moving upward,
431
opposite the direction of the ligament and most of the surrounding slurry. This upward moving steam is likely
432
helping strip away droplets via shear. By the time the wave has collapsed (frame 6), a clear divide is observed
433
around the middle of the beach. To the left, slurry is moving upwards, and to the right, slurry is moving
434
downwards. The primary exception to this trend is the downward moving slurry in the sheltered region on the
435
leeward side of the newly formed wave.
436
437
438
19
439
Figure 9 The ratio of radial velocity to the absolute value of axial velocity through one representative wave
440
cycle. Contours are on a plane corresponding to a 30° azimuthal angle. As the wave peaks in frame 4, much
441
of the wave base is moving in the axial direction (white). The wave tip and associated ligaments, however, are
442
still slinging upward (negative radial direction) into the steam flow, which will enhance droplet production.
443
444
Figure 10 shows contours at the moment in time roughly where the wave peaks in height (equivalent to frame
445
4 in Figure 9) at 6 azimuthal angles. The base of the wave is generally uniform across the various azimuthal
446
slices, but the wave tips differ significantly across angles, even at 6° increments. What is consistent across
447
angles is this: while the wave base is moving with a largely axial velocity, the wave tip has a significant radial
448
velocity component. In frames 1 and 6 especially, we see a distinct secondary ligament form behind the wave
449
tip, which also has a high radial velocity. Most frames include some level of a steam gap between the wave tip
450
and a secondary ligament or wave base. The resulting wave tips can take on something of a hammerhead
451
shape before breaking away (for example, frames 2, 4, and 6) as the slurry connecting the tip to the base is
452
thinned. We direct the reader to notice particularly the penetration of steam into the wave as the tip rises. In
453
frames 2-4, the steam is pushing down into the wave, creating significant necking to break off the tip. Frame
454
5 shows smaller pockets of steam inside the wave, where steam fingers have forced through the deformed
455
surface. The upward movement of the wave, then, provides an effervescence mechanism in addition to the
456
trapping of steam as the wave crashes.
457
458
459
460
20
Figure 10 The ratio of radial velocity to the absolute value of axial velocity at a fixed time, roughly where the
461
wave peaks (equivalent to frame 4 in Figure 9). Contours are on planes corresponding to azimuthal angles of
462
30°, 36°, 42°, 48°, 54°, and 60°. The wave base consistently moves largely in the axial direction (white), while
463
the wave tip generally includes a significant upward velocity component. Steam penetrates the wave to break
464
off the tip (frames 2-4) and as smaller bubbles (frame 5), providing a mechanism for effervescence.
465
(Multimedia view).
466
467
It is clear that flicking is an important mechanism for droplet production in the nozzle, but is stripping also
468
important? While flicking involves ligaments rising perpendicular to the steam flow, by stripping, we mean
469
droplets breaking away from the slurry surface or a ligament stretching in the direction of steam flow (parallel).
470
A close look at Figure 9 seems to show stripping towards the beginning of wave growth, but this is made
471
clearer in the animation for Figure 10. This animation shows stripping early on where droplets break free from
472
both the surface and parallel ligaments. As the wave matures, the primary droplet production mechanism
473
transitions to flicking. Finally, outside the nozzle, effervescence and the collapse of the wave lead to its
474
complete disintegration. In summary, droplet production inside the nozzle is dominated by stripping in the
475
early wave and later by flicking as the wave crests.
476
477
To conclude our discussion, we note several consequences (and benefits) of slurry ligaments flicking radially
478
up into the gas stream. Effervescence is introduced via a second mechanism before the wave crashes and
479
sandwiches steam. While the slurry wave jet is accelerated by the steam and thinned by shear, causing the jet
480
to buckle upwards and sling the liquid tips25 up into the gas stream as ligaments, the gas infiltrates the liquid
481
as bubbles. The thin fingers from the liquid sheet make it easier for the steam to peel off droplets,43 and this
482
avoids the need for a forcibly thinned sheet using a thin slurry annulus with high pressure drop and risk of
483
plugging. Additionally, the shearing and thinning of the fingers lowers the timescale for RTI to take effect,
484
making RTI more active and sooner.
485
486
3.3 Point Monitors
487
3.3.1 Overview
488
In a previous study, only two quantities were tracked at a single point monitor.22 We greatly expand point
489
monitor analysis here to include 100 signals for the purpose of understanding the spatial and temporal variation
490
21
of quantities, both azimuthally and along the slurry flow path. Velocity magnitude, slurry volume fraction
491
(VF), strain rate, and turbulent kinetic energy (TKE) were tracked at five point monitors placed roughly along
492
the trajectory of the slurry interface as it interacts with the steam. Each point monitor is present at five different
493
azimuthal angles (4.5°, 13.5°, 22.5°, 31.5°, and 40.5°), spaced evenly within a 45° azimuth, which is half of
494
the total 90° azimuth. In total then, 100 individual signals were recorded (4 quantities across 25 points) for the
495
entirety of the QSS window. Due to the extensive nature of this temporal data collection, we provide only
496
select plots and summary statistics. The locations of the five point monitors in a given azimuthal slice are
497
shown in Figure 11. The labels of Inner, Middle, Exit, Tip, and Outer, will continue to be used in reference to
498
these points, along with their azimuthal angles. The Inner, Middle, Exit, Tip, and Outer points are located 1,
499
0.75, 0.67, 1.25, and 1.5 nozzle radii from the nozzle axis (bottom edge of Figure 11), respectively. The Inner
500
point is centered axially on the wave pool. The Middle and Exit points are positioned axially at the start and
501
end of the beach. The Tip and Outer points are located 0.25 and 1.25 nozzle radii downstream of the nozzle
502
exit, respectively. The output data monitored at these five points will be discussed throughout the remainder
503
of the paper, revealing the spatiotemporal characteristics of the system.
504
505
506
507
Figure 11 Locations of five points monitors (roughly along the slurry interface trajectory) where certain data
508
quantities are collected. Each point monitor exists at 5 azimuthal angles within a 45° azimuth (half of the total
509
90° azimuth): 4.5°, 13.5°, 22.5°, 31.5°, and 40.5°.
510
511
Comparing the Exit point monitors across azimuthal angles reveals aspects of the wave interface as it exits the
512
nozzle. Figure 12 shows a moving average time series for (a) velocity and (b) slurry volume fraction through
513



 
22
12.5 pulsing cycles. Velocity signals are fairly uniform across angles and from wave to wave. In other words,
514
the general motion of the wave seems to vary minimally as it exits the nozzle, both azimuthally and temporally.
515
However, slurry volume varies significantly in the azimuthal and temporal dimensions. We attribute this
516
variability largely to the work of RTI, which is exacerbated by windward pressure buildup. We note again that
517
it is around the nozzle exit that the wave collapses, marking a transition from a somewhat 2D wave to a more
518
azimuthally diverse 3D wave. Table I provides the mean and coefficient of variation (COV) for each signal in
519
Figure 12 as well as the other quantities that are not displayed graphically. Note that this is a small subset of
520
all point monitor data. The COV is a normalized standard deviation, where the standard deviation is divided
521
by the mean and converted to a percentage.
522
523
524
Figure 12 Moving average time series of (a) velocity and (b) slurry volume fraction at the Exit point monitors
525
across all five azimuthal angles. The velocity signals are fairly uniform, both azimuthally and temporally. RTI
526
causes slurry volume fraction, however, to vary significantly in the azimuthal and temporal dimensions.
527
528
529
Table I Mean and coefficient of variation (COV) for signals across all five azimuthal angles at the Exit point.
530
COV is calculated by dividing the standard deviation by the mean and converting to a percentage.
531
532
Velocity [m/s]
Slurry VF
Strain Rate [1/s]
TKE [m2/s2]
Angle
COV
COV
Mean
COV
Mean
COV
4.5
94
219
4.5×105
156
412
142
13.5
83
312
4.5×105
141
414
129
22.5
83
311
3.6×105
152
345
131
31.5
82
301
3.0×105
170
325
144
40.5
79
299
2.9×105
166
312
148
533
23
On the other hand, we can reveal the axial (roughly) development of quantities by examining all five point
534
monitors at a given azimuthal angle. Figure 13 shows the moving average time series for (a) velocity and (b)
535
strain rate for all point monitors at the 22.5° azimuthal angle. Velocity magnitudes increase significantly as
536
the wave moves from the pool to the nozzle exit but decreases at the Tip and Outer point monitors. The slurry
537
is disintegrating at these last two points, so the monitors are picking up both slurry droplets and relatively
538
stagnant steam. Periodicity is much more evident inside the nozzle than outside the nozzle. This is indicative
539
of the relatively uniform wave motion inside the nozzle and the chaotic rupture outside the nozzle. Strain rate
540
follows a similar pattern through the slurry motion, increasing to the nozzle exit and then decreasing outside
541
the nozzle. However, the difference between the Middle/Exit points and the Tip/Outer points is less marked
542
than velocity. Compared to velocity, the temporal periodicity of strain rate is less pronounced for the middle
543
Middle/Exit points and more pronounced for the Tip/Outer points. In other words, velocity gradients are
544
fluctuating more consistently than velocity magnitude outside the nozzle. Strain rate also shows more wave-
545
to-wave variation than velocity within the nozzle. Table II provides the mean and COV for each signal in
546
Figure 13 as well as the other quantities that are not displayed graphically.
547
548
549
Figure 13 Moving average time series of (a) velocity and (b) strain rate at all five point monitors at the 22.5°
550
angle, illustrating variation of these quantities spatiotemporally as slurry moves through the system. The
551
velocity increases up to the nozzle exit as waves form and is much more periodic inside the nozzle than outside
552
the nozzle, showing the contrast between the more ordered wave formation and the more chaotic radial
553
bursting. Inside the nozzle, strain rate shows more wave-to-wave variation than velocity, and the Tip and Outer
554
points show more periodicity.
555
556
557
558
24
Table II Mean and coefficient of variation (COV) for signals across all point monitors at the 22.5° azimuthal
559
angle. COV is calculated by dividing the standard deviation by the mean and converting to a percentage.
560
561
Velocity [m/s]
Slurry VF
Strain Rate [1/s]
TKE [m2/s2]
Point
COV
COV
Mean
COV
Mean
COV
Inner
18
0.0
1.9×103
54
1.1
60
Middle
86
176
2.2×105
136
255
147
Exit
83
311
3.6×105
152
345
131
Tip
67
367
6.8×104
257
28
326
Outer
46
550
7.7×104
137
14
320
562
563
3.3.2 Frequency Analysis
564
FFTs were performed to determine the dominate frequencies in the atomizer at various locations. Because of
565
the pulsing nature of the system, we expect most quantities to cycle at a consistent overall pulsing frequency
566
(frequency of wave formation). Velocity magnitude, strain rate, slurry volume fraction, and turbulent kinetic
567
energy were tracked at all 25 point monitor locations. FFTs reveal frequencies around 1000 Hz for the vast
568
majority of these quantities and point monitors. 83% of signals show a 1068 Hz peak frequency (note that
569
slurry volume does not fluctuate at any Inner point locations). 8% show a 2060 Hz peak frequency, and one
570
shows 3128 Hz, illustrating the higher mode harmonics. Strain rate is the most consistent quantity: all 25 point
571
monitors show a dominate frequency of 1068 Hz. The most prominent frequencies across point monitors for
572
the Ref-1 and Ref-2 meshes are 963 Hz and 992 Hz, respectively. Frequencies vary wildly for the Base mesh,
573
and no prominent frequency is evident. This corresponds to previous findings: major wave characteristics are
574
present, and consistent, with progressively increased mesh resolution, beginning with the resolution of Ref-
575
1.22 The characteristic pulsing of the system is largely absent from the Base case, and the Ref-3 mesh is two
576
refinement levels above Ref-1.
577
578
To understand how FFT peak frequencies vary across the azimuth, the COV was computed, which shows
579
azimuthal point-to-point variation as a percentage. For a given point location, such as Inner, a single peak
580
frequency was computed at each azimuthal angle, and the COV was computed from these 5 frequencies. Figure
581
14 shows the azimuthal frequency COV from the Inner to the Outer points (effectively along the interfacial
582
trajectory). Most frequencies are consistent inside the nozzle, but the variation increases significantly outside
583
25
the nozzle. The only quantity that continues to fluctuate uniformly in the azimuth outside the nozzle is strain
584
rate. Figure 14 illustrates again the consistency of strain rate as a pulsating quantity.
585
586
587
Figure 14 Azimuthal coefficient of variation (COV) for frequency at various points along the slurry interface
588
flow path. COV represents the percent of variation in the azimuth at each point. VF is volume fraction, and
589
TKE is turbulent kinetic energy. The peak frequencies were determined by FFTs. Frequencies are generally
590
less consistent outside of the nozzle as the slurry disintegrates. Strain rate stands out, fluctuating at consistent
591
frequencies at all locations (inside the nozzle, outside the nozzle, and across the azimuth).
592
593
FFTs are meaningless unless the resulting peak frequencies have converged over the course of the simulation.
594
For those signals that showed a peak frequency of 1068 Hz (the prominent frequency among all signals
595
evaluated), the FFTs generally converged within 4 pulsing cycles after the flow was already at QSS. This was
596
not necessarily the case for less periodic signals like velocity at the Tip and Outer points (see Figure 13).
597
Figure 15 provides a sample FFT peak frequency convergence plot for strain rate at the Exit point in the 22.5°
598
plane (see Figure 13 for time series). The vertical dashed line is the 4 pulsing cycles mark, and the green star
599
is the final value. Zero-padding was employed, resulting in the discrete step values towards convergence.
600
601







    






26
602
603
Figure 15 Convergence of FFT peak frequency for the strain rate signal at the Exit point monitor in the 22.5°
604
plane. FFTs included zero-padding, which produces the discrete steps towards convergence. The vertical
605
dashed line is at 4 pulsing times, and the green star represents the final frequency value of 1068 Hz. Signals
606
generally converged to this value within 4 pulsing cycles.
607
608
Figure 16 presents two FFT examples, one for velocity at the Inner point (left) and one for strain rate at the
609
Exit point (right). Both are at a 22.5° azimuthal angle, and the time series were shown in Figure 13. The Inner
610
point is located at the surface of the wave pool, where waves are being produced. The wave pool surface, then,
611
has a fluctuating velocity at 1068 Hz, which sets the pace for bulk pulsation in the system. Geometric
612
parameters might be varied to determine their influence on pulsing frequency, but that is beyond the scope of
613
this study. The Inner velocity FFT also shows harmonics at roughly 2060 Hz and 3110 Hz, which correspond
614
to the peak frequencies for a minority of signals. The Exit strain rate FFT shows the same peak frequency as
615
the Inner velocity, although it is less pronounced, and no harmonics are evident.
616
617
618
27
Figure 16 FFTs with peak frequency labeled for (a) velocity at the Inner point and (b) strain rate at the Exit
619
point. Both points are at an azimuthal angle of 22.5°. The velocity at the Inner point (where waves are formed)
620
shows a peak frequency of 1068 Hz and clear harmonics at around 2060 Hz and 3110 Hz. Strain rate at the
621
Exit point has the same peak frequency, but it is less distinct and with no harmonics.
622
623
As shown, the preponderance of frequencies are close to, or multiples of, approximately 1000 Hz, which results
624
from the wave pool generation process (i.e. KHI working with Bernoulli to amplify surface disturbances).
625
Therefore, the wave-generation process sets up an absolute instability in the system that is likely to be
626
unaffected by upstream turbulence effects. A future study could include evaluations of this.
627
628
3.3.3 Cross-Correlations
629
The spread of information in the azimuthal and radial directions can be assessed by determining the cross-
630
correlation between signals; a normalized cross-correlation of 1 indicates that two time-series signals are
631
perfectly correlated and implies 2D (axisymmetric) motion. The time difference between two given locations
632
was accounted for by time-shifting the signals to align them before calculating the normalized cross-
633
correlation. Cross-correlations (calculated using the NumPy package in Python) were normalized by
634
subtracting the means from the signals and dividing the cross-correlation by the number of data points and the
635
standard deviations of the two signals. Figure 17 shows the normalized cross-correlation between velocity
636
signals both azimuthally and along the slurry interface trajectory. Correlation is calculated between the velocity
637
signal at a given azimuthal angle or point and the first angle or point. In essence, we are estimating how much
638
the motion at one place in the flow field might be related to the motion at another place in the flow field.
639
640
The motion at the inner point monitor appears to be uniform across all angles (indicating its 2D nature), as the
641
signals show almost perfect correlation with the first angle (4.5°). The Middle and Exit points show relatively
642
high correlation, and the correlation remains largely the same across angles. Outside of the nozzle, where the
643
wave and slurry sheet are being ruptured, the tip and outer point monitors show low correlation between angles.
644
The Tip point monitor, which is closer to the nozzle (around the bulging and bursting), shows slightly higher
645
correlation. These results suggest that fluid motion inside the nozzle is generally azimuthally similar (2D), but
646
the fluid motion outside the nozzle as the slurry bursts is azimuthally unrelated (3D). Furthermore, since the
647
28
correlation, for a given point monitor, is quite consistent across 45°, we conclude that a 90° mesh is more than
648
sufficient to capture azimuthal variation. The Inner point, which is at the surface of the wave pool, marks the
649
location of wave generation. The Middle and Exit points are well-correlated with the Inner point, but the Tip
650
and Outer points are not. In other words, the velocities at Tip and Outer do not show much relation to wave
651
pool motion.
652
653
654
655
Figure 17 Normalized cross-correlation between transient velocity magnitude signals from point monitors at
656
five points along the slurry interfacial flow and five azimuthal angles. A given normalized cross-correlation
657
was calculated after the two signals were time-shifted to align. A value of 1 indicates perfect correlation
658
between signals. The correlation between a given angle or point with the first angle (4.5°) or point (Inner) is
659
computed. High correlation is observed both azimuthally and along the slurry flow inside the nozzle but is
660
greatly diminished outside the nozzle. The lack of azimuthal variation in correlation across 45° demonstrates
661
the sufficiency of a 90° mesh to capture variations in the azimuth.
662
663
When signals are reasonably correlated, a time lag shows how much the motions are temporally offset. The
664
time lag is calculated as the amount the signals must be time-shifted to produce the maximum cross-correlation.
665
In Figure 18, time lags are normalized by the pulsing time, so a given value represents the fraction of a pulse
666
cycle by which the signals are offset. Time lag is meaningless if the signals are not well-correlated, so the Tip
667
and Outer points have been removed from the time lag plot. Figure 18 displays the extent to which the Middle
668
and Exit velocity signals lag the Inner velocity signal. These results are communicating where the Middle and
669
Exit points are, temporally, within the wave cycle (which repeats regularly every 1 PT). Velocity fluctuations
670
at the wave pool (Inner point) reach the Middle and Exit points 0.7 and 0.8 PTs, respectively, after the initial
671
wave pool motion. Fluid motion at the nozzle exit then lags the Middle point by 0.1 PT. Velocity spikes again
672
29
in the wave pool 0.2 PTs after a velocity increase at the Exit point. This cycle is illustrated in Figure 18, which
673
shows the percentage of a pulse cycle for velocity fluctuations at Inner (I) to reach Middle (M) and then Exit
674
(E) and then start again at Inner. The majority of a given pulse cycle (70%) involves the wave growing out of
675
the wave pool and reaching the beach. After reaching the beach, the wave travels more rapidly to the nozzle
676
exit and beyond.
677
678
Figure 18 Transient velocity signal time lags from the Inner point at five azimuthal angles. The time lag
679
corresponds to the signal offset that produces the maximum cross-correlation and has been normalized by the
680
pulsing time. Wave pool velocity fluctuations (Inner point, I) take 0.7 and 0.8 pulsing times to reach the Middle
681
(M) and Exit (E) points, respectively. Initial growth of a wave consumes the majority (70%) of a given pulse
682
cycle.
683
684
In addition to inter-point correlations for velocity, correlations between different quantities were calculated at
685
all 25 points. Three contour plots in Figure 19 summarize these data and show the normalized cross-correlation
686
between 1) velocity magnitude and slurry VF (far left), 2) TKE and strain rate (middle), and 3) velocity
687
magnitude and strain rate (far right). The slurry volume fraction maintains a value of 1 at the Inner points,
688
making any normalized cross-correlation value meaningless. For this reason, the inner points were excluded
689
on the leftmost plot in Figure 19. The trend across all 3 sets of correlations is that any two quantities are most
690
highly correlated at the Inner point, and correlation decreases downstream. Correlation between quantities is
691
also fairly azimuthally uniform, particularly inside the nozzle. Velocity and slurry VF are the least well-
692
correlated overall. The middle plot indicates that, after the Exit point, TKE and strain rate are decoupled; thus,
693
TKE downstream must have been produced by some earlier shear. In summary, we observe the decoupling of
694
quantities exiting the nozzle: strong correlations inside the nozzle and very weak correlations outside the
695
nozzle.
696
30
697
698
Figure 19 Contours across all 25 points for the normalized cross-correlations between 1) velocity magnitude
699
and slurry volume fraction (left), 2) turbulent kinetic energy strain rate (middle), and 3) velocity magnitude
700
and strain rate (right). The Inner points have been excluded on the leftmost plot because the slurry volume
701
fraction maintains a value of 1. Correlations are generally strong inside the nozzle but very weak outside the
702
nozzle, showing a decoupling of quantities exiting the nozzle.
703
704
705
3.4 Adaptive Mesh Refinement
706
Up to this point, the Ref-3 mesh has been used entirely for visualization and analysis. It has been noted that
707
Ref-3-AMR, though a more cost-efficient alternative, is the outlier to the numerical trends of two point monitor
708
signals,22 but no visual comparison was provided, and a more rigorous comparison is lacking. A qualitative
709
and quantitative comparison of Ref-3 and Ref-3-AMR is here provided to clarify the differences between the
710
two models. We emphasize again that Ref-3 and Ref-3-AMR have the exact same mesh resolution at the slurry-
711
steam interface. A rigorous VOF gradient criterion and adaption frequency were used, but it should be noted that
712
there are other criteria that can determine the regions of adaption for AMR. It is possible that a different criterion
713
would positively affect AMR results, though we find this doubtful based on the extensive AMR refinement in this
714
study. We also acknowledge that any conclusions about AMR as a technique are limited to ANSYS Fluent and its
715
underlying algorithms. All validation efforts were conducted using meshes equivalent to Ref-3 (mostly
716
hexahedral elements swept in the flow direction), and AMR has not been validated experimentally for this
717
work. The differences between Ref-3 and Ref-3 AMR indicate problems with the communication of
718
information through split cells, which are constantly created by AMR.
719
720
Figure 20, which shows the slurry surface (coloured by slurry viscosity) as it exits the nozzle (flow is generally
721
downward), reveals qualitative differences. The leftmost set of images in Figure 20 show that Ref-3-AMR
722
31
produces a significantly more rippled surface, and slurry viscosity is generally higher. A higher viscosity
723
indicates lower strain rate, perhaps suggesting that the mesh gradients around the slurry-steam interface are
724
affecting velocity gradients. We remind the reader that RTI and viscosity gradients serve to destabilize the
725
annular slurry sheet, priming it for rupture and atomization as the wave crashes into it. The middle and
726
rightmost images in Figure 20 reveal Ref-3 rupturing more readily than Ref-3-AMR. In the middle images,
727
little pre-rupture (localized bursting events) in the annular slurry sheet is observed for Ref-3-AMR. Much more
728
can be seen for Ref-3 (corresponding to the preliminary rupture for Ref-3 in frame 5 of Figure 4). In the
729
rightmost images, the radial burst of slurry is more violent for Ref-3 than Ref-3-AMR, and Ref-3 is clearly
730
producing smaller droplets at this stage of the pulsing sequence. Interestingly, Ref-3-AMR was found to have
731
slightly larger droplets throughout the domain past the nozzle exit.23
732
733
Quantitative discrepancies between Ref-3 and Ref-3-AMR are summarized in Figure 21, which shows percent
734
differences between azimuthally-averaged quantities. TKE shows by far the greatest difference between the
735
models, especially outside the nozzle, where a 160% difference is observed. In general, the difference between
736
model outputs is higher outside (shown in Figure 20) than inside the nozzle. This follows the trend of an
737
increasing presence of mesh element size gradients as the slurry disintegrates outside the nozzle. Interestingly,
738
the % difference for strain rate (velocity gradients) increases outside the nozzle, but that for velocity does not.
739
This observation indicates that gradients are more strongly affected by AMR. We note also that TKE
740
production is driven by velocity gradients. Our conclusion: the AMR technique within ANSYS Fluent 2020R1
741
is not sufficient to accurately model non-Newtonian wave-augmented atomization in the present system.
742
743
32
744
745
Figure 20 Comparison of slurry surface as computed by the Ref-3 (top row) and Ref-3-AMR (bottom row)
746
meshes at three points in the wave cycle. Both models represent a 90° azimuthal slice with periodic boundary
747
conditions. The top three images are at roughly the same stages within a given pulsing cycle as the bottom
748
three images. Ref-3-AMR produces a noticeably more rippled surface than Ref-3. In the Ref-3-AMR case, the
749
slurry sheet stretching down from the nozzle maintains a higher viscosity and does not rupture as quickly as
750
Ref-3.
751
752
753
Figure 21 The percent difference between azimuthally-averaged quantities at various points along the slurry
754
interface flow path for Ref-3 and Ref-3-AMR. VF is volume fraction, and TKE is turbulent kinetic energy.
755
TKE shows the greatest difference between the models, with % differences above 150% outside the nozzle.
756
The general trend is a greater divergence in model outputs outside than inside the nozzle.
757
758
759
760









    






33
4. CONCLUSION
761
Using a computational framework which has been validated and used extensively over the course of the last
762
decade, we have analyzed the spatiotemporal characteristics of the novel WAVE process (Wave-Augmented
763
Varicose Explosions) in which non-Newtonian waves facilitate disintegration in an inverted feed twin-fluid
764
atomizer. Annular slurry hot steam flow, creating a secondary nozzle effect for the
765
steam, as an annular slurry sheet stretches from the nozzle. Waves then collapse as they exit the nozzle,
766
crashing into the slurry sheet in a violent radial burst to enhance droplet formation. The wave birth-death cycle
767
is part of a general bulk system pulsation phenomenon, causing many quantities to fluctuate periodically.
768
Important knowledge gaps from previous studies have been filled to provide a complete understanding of this
769
efficient atomization phenomenon. A summary of new contributions to the literature is provided in the
770
following paragraphs.
771
772
We began by presenting a unique visual perspective: unraveled wave views to elucidate characteristics of the
773
3D wave cycle. An estimate of RTI time scale showed sufficient time for RTI development, and baroclinic
774
torque on the order of  1/s2 indicates strong RTI activity in the wave. KHI and RTI cause surface
775
variations in the non-Newtonian slurry that then excite the instabilities in a self-amplification cycle. During
776
the early stages of wave growth, stripping is a dominant mechanism for droplet production. Later, as the wave
777
crests, ligaments flicking up into the steam flow at the wave tip facilitate droplet production inside the nozzle.
778
Meanwhile, the radial thrust of the wave allows for steam penetration to increase effervescence and sometimes
779
break off the wave tip. Further disintegration is encouraged as the wave leaves residual ligaments in its wake.
780
781
An evaluation of velocity, slurry volume fraction, strain rate, and turbulent kinetic energy at 25 axially and
782
azimuthally spaced points revealed a loss of consistent fluctuation frequency outside of the nozzle. Strain rate,
783
however, cycles with the dominant system frequency at all points. FFT analysis was an important component
784
of this study, and peak frequencies generally converged within 4 pulsing cycles. The nozzle exit marks a
785
significant increase in azimuthal variation of velocities. Velocities show strong azimuthal correlation (2D) in
786
the wave formation region inside the nozzle but are azimuthally unrelated (3D) outside the nozzle, where radial
787
34
bursting is occurring. Outside the nozzle, fluid motion did not show strong correlation with wave pool motion.
788
Correlations between quantities, though strong in the wave formation region, showed a consistent trend of
789
significant decoupling outside the nozzle. Time lags revealed that, for a given pulse cycle, the wave rising out
790
of the pool to reach the beach takes 70% of the total pulse time.
791
792
Azimuthal correlations of velocity demonstrate that a 90° angle is sufficient for to capture azimuthal
793
variability, which is important for atomization. Evaluation of Ref-3-AMR, both qualitatively and
794
quantitatively, showed that this particular AMR technique is insufficient for accurately modeling non-
795
Newtonian, wave-augmented atomization in the present system, despite the allure of computational savings.
796
Velocity gradients were more affected by AMR than velocity magnitude. Turbulent kinetic energy differed
797
most drastically between Ref-3 and Ref-3-AMR, particularly outside the nozzle. Future research could
798
evaluate how geometry changes affect the spatiotemporal characteristics of the atomization process, such as
799
pulsing frequency.
800
801
ACKNOWLEDGEMENT
802
The authors thank Valda Rowe and Dr. Mark Horstemeyer for their administrative support.
803
804
DATA AVAILABILITY
805
The data that support the findings of this study are available from the corresponding author upon reasonable
806
request.
807
808
CONFLICT OF INTEREST
809
The authors have no conflicts to disclose.
810
811
REFERENCES
812
1 Abhijeet Kumar and Srikrishna Sahu, "Large scale instabilities in coaxial air-water jets with annular air swirl," Physics of
813
Fluids 31 (12), 124103 (2019).
814
35
2 Kiumars Khani Aminjan et al., "Study of pressure-swirl atomizer with spiral path at design point and outside of design
815
point," Physics of Fluids 33 (9), 093305 (2021).
816
3 Shirin Patil and Srikrishna Sahu, "Air swirl effect on spray characteristics and droplet dispersion in a twin-jet crossflow
817
airblast injector," Physics of Fluids 33 (7), 073314 (2021).
818
4 Erkki Laurila et al., "Numerical study of bubbly flow in a swirl atomizer," Physics of Fluids 32 (12), 122104 (2020).
819
5 Mayukhmali Chakraborty, Aravind Vaidyanathan and S. L. N. Desikan, "Experiments on atomization and spray
820
characteristics of an effervescent strut injector," Physics of Fluids 33 (1), 017103 (2021).
821
6 Georgios Charalampous, Constantinos Hadjiyiannis and Yannis Hardalupas, "Proper orthogonal decomposition of primary
822
breakup and spray in co-axial airblast atomizers," Physics of Fluids 31 (4), 043304 (2019).
823
7 Hao (武浩) Wu, Fujun (张付军) Zhang and Zhenyu (章振宇) Zhang, "Droplet breakup and coalescence of an internal-
824
mixing twin-fluid spray," Physics of Fluids 33 (1), 013317 (2021).
825
8 Abhijeet Kumar and Srikrishna Sahu, "Liquid jet disintegration memory effect on downstream spray fluctuations in a coaxial
826
twin-fluid injector," Physics of Fluids 32 (7), 073302 (2020).
827
9 Santanu Kumar Sahoo and Hrishikesh Gadgil, "Dynamics of Self-Pulsation in Gas-Centered Swirl Coaxial Injector: An
828
Experimental Study," Journal of Propulsion and Power 37 (3), 450-462 (2021).
829
10 Santanu Kumar Sahoo and Hrishikesh Gadgil, "Large scale unsteadiness during self-pulsation regime in a swirl coaxial
830
injector and its influence on the downstream spray statistics," International journal of multiphase flow 149, 103944 (2022).
831
11 Sergey V. Alekseenko et al., "Study of formation and development of disturbance waves in annular gasliquid flow,"
832
International Journal of Multiphase Flow 77, 65-75 (2015).
833
12 N. J. Balmforth, R. V. Craster and C. Toniolo, "Interfacial instability in non-Newtonian fluid layers," Physics of Fluids 15
834
(11), 3370-3384 (2003).
835
13 S. Millet et al., "Wave celerity on a shear-thinning fluid film flowing down an incline," Physics of Fluids 20 (3), 031701
836
(2008).
837
14 Hui Zhao et al., "Secondary breakup of coal water slurry drops," Physics of Fluids 23 (11), 113101 (2011).
838
36
15 Jin-Peng (郭瑾朋) Guo et al., "Instability breakup model of power-law fuel annular jets in slight multiple airflows," Physics
839
of Fluids 32 (9), 094109 (2020).
840
16 Wayne Strasser, "Towards Atomization for Green Energy: Viscous Slurry Core Disruption by Feed Inversion," AAS 31 (6),
841
23-43 (2021).
842
17 Daniel M. Wilson and Wayne Strasser, "Smart Atomization: Implementation of PID Control in Biosludge Atomizer," 5-6th
843
Thermal and Fluids Engineering Conference , (2021).
844
18 Manisha B. Padwal and D. P. Mishra, "Interactions among synthesis, rheology, and atomization of a gelled propellant,"
845
Rheol Acta 55 (3), 177-186 (2016).
846
19 Manisha B. Padwal and D. P. Mishra, "Effect of Air Injection Configuration on the Atomization of Gelled Jet A1 Fuel in an
847
Air-Assist Internally Mixed Atomizer ," Atomization and Sprays 23 (4), 327-341 (2013).
848
20 Manisha B. Padwal and Debi Prasad Mishra, "Performance of Two-Fluid Atomization of Gel Propellant," Journal of
849
propulsion and power 38 (1), 30-39 (2022).
850
21 Li-jun Yang et al., "Linear Stability Analysis of a Non-Newtonian Liquid Sheet," Journal of propulsion and power 26 (6),
851
1212-1225 (2010).
852
22 D. M. Wilson and W. Strasser, "The Rise and Fall of Banana Puree: Non-Newtonian Annular Wave Cycle in Transonic
853
Self-Pulsating Flow," Physics of Fluids 34 (7), (2022).
854
23 D. M. Wilson and W. Strasser, "A Spray of Puree: Wave-Augmented Transonic Airblast Non-Newtonian Atomization,"
855
Physics of Fluids 34 (7), (2022).
856
24 Cynthia Ditchfield et al., "Rheological Properties of Banana Puree at High Temperatures," International Journal of Food
857
Properties 7 (3), 571-584 (2004).
858
25 W. Mostert, S. Popinet and L. Deike, "High-resolution direct simulation of deep water breaking waves: transition to
859
turbulence, bubbles and droplets production," Journal of fluid mechanics 942 (A27), (2022).
860
26 Chenwei (张宸玮) Zhang et al., "Atomization of misaligned impinging liquid jets," Physics of Fluids 33 (9), 093311
861
(2021).
862
37
27 Wayne Strasser and Francine Battaglia, "The effects of pulsation and retraction on non-Newtonian flows in three-stream
863
injector atomization systems," Chemical Engineering Journal 309, 532-544 (2017).
864
28 David Youngs, Time-Dependent Multi-material Flow with Large Fluid Distortion, in Numerical Methods in Fluid
865
Dynamics, edited by K. W. Morton and M. J. Baines, (Academic Press, 1982), pp. 273-285.
866
29 Suhas V. Patankar, Numerical Heat Transfer and Fluid Flow, edited by Anonymous 1st ed. (CRC Press, London, 1980), .
867
30 Subramanian Easwaran Iyer et al, US Patent No. 8734909 (May 27, 2014).
868
31 W. Strasser and F. Battaglia, "Pulsating Slurry Atomization, Film Thickness, and Azimuthal Instabilities," Atomization and
869
Sprays 28 (7), 643-672 (2018).
870
32 Wayne Strasser, "Towards the optimization of a pulsatile three-stream coaxial airblast injector," International Journal of
871
Multiphase Flow 37 (7), 831-844 (2011).
872
33 Wayne Strasser and Francine Battaglia, "Identification of Pulsation Mechanism in a Transonic Three-Stream Airblast
873
Injector," Journal of Fluids Engineering 138 (11), (2016).
874
34 Wayne Strasser and Francine Battaglia, "The Influence of Retraction on Three-Stream Injector Pulsatile Atomization for
875
AirWater Systems," Journal of Fluids Engineering 138 (11), (2016).
876
35 Wayne Strasser, Francine Battaglia and Keith Walters, "Application of a Hybrid RANS-LES CFD Methodology to Primary
877
Atomization in a Coaxial Injector," Volume 7A: Fluids Engineering Systems and Technologies , (2015).
878
36 Wayne Strasser and Francine Battaglia, "The Effects of Prefilming Length and Feed Rate on Compressible Flow in a Self-
879
Pulsating Injector ," AAS 27 (11), (2017).
880
37 Wayne Strasser, "Oxidation-assisted pulsating three-stream non-Newtonian slurry atomization for energy production,"
881
Chemical Engineering Science 196, 214-224 (2019).
882
38 Wayne Strasser, "The war on liquids: Disintegration and reaction by enhanced pulsed blasting," Chemical Engineering
883
Science 216, 115458 (2020).
884
39 Reid Prichard and Wayne Strasser, "Optimizing Selection and Allocation of High-Performance Computing Resources for
885
Computational Fluid Dynamics," 7th Thermal and Fluids Engineering Conference (under review) , (2022).
886
38
40 Wayne Strasser, "Toward Atomization for Green Energy: Viscous Slurry Core Disruption by Feed Inversion," AAS 31 (6),
887
(2021).
888

889
Taylor Instabilities," Journal of geophysical research. Solid earth 123 (5), 3593-3607 (2018).
890
42 P. K. Senecal et al., "Modeling high-speed viscous liquid sheet atomization," International journal of multiphase flow 25
891
(6), 1073-1097 (1999).
892
43 Yue Ling et al., "Spray formation in a quasiplanar gas-liquid mixing layer at moderate density ratios: A numerical closeup,"
893
Physical review fluids 2 (1), (2017).
894
895
896
... Given the low , the model was treated as isothermal at ambient conditions to match the biofuel 205 experiment. These methods have been extensively validated and used in the past [7,21,[47][48][49][50] . ...
... In contrast to pressure atomization which relies on considerable injection pressure, effervescent atomization is less dependent on the injecting pressure and therefore has been widely studied. 10 Generally, twin-fluid injectors can be categorized into internal and external mixing types depending on the mixing mode. 11 As for the internal mixing mode, there are two means of introducing atomizing gas medium into the liquid stream inside the injector: outside-in-gas (OIG) and outside-in-liquid (OIL) types. ...
Article
Full-text available
Passive control of twin-fluid atomization can be achieved by changing the orifice shape of the injector. In this study, the characteristics of twin-fluid atomization in the outside-in-liquid injector with circular, square, and rectangular orifices at various aspect ratios were investigated experimentally and computationally. The morphology of the spray was captured by shadowgraph, the diameter and velocity of the droplets were measured by the phase Doppler particle analyzer, and numerical simulations were performed for the central gaseous core. Comparing the sprays with square and circular orifices, droplets from the non-circular orifice are generally smaller with less disparities in droplet sizes due to the more intensive turbulent disturbances and corner effect. Furthermore, the non-circular orifice also results in better spatial distribution of the spray. The droplet diameters of the spray with a square orifice do not satisfy the log-normal distribution near the orifice along the centerline of the spray, which may be attributed to the different entrainment of spray droplets by the central gas flow for the sprays with circular and non-circular orifices. The twin-fluid sprays produced by the rectangular orifice also exhibit the same axial switching effect as in the high-pressure gaseous jet flow, in which the spray diffusion in the minor axis is more extensive than that in the major axis. Moreover, the droplets' Sauter mean diameter produced by the rectangular orifice is more sensitive to the size in the minor axis of the orifice and decreases as the aspect ratio of the orifice increases given the same cross-sectional area.
Article
This paper mainly reviews the research progress on multiphase flow in nuclear engineering. This paper is composed of three parts. The first part documents a literature statistical analysis based on publications, mainly the SCIE-indexed papers collected in the Web of Science. The analysis data of publications on multiphase flows showcase the prosperous future of the current Experimental and Computational Multiphase Flow journal. The second part summarizes advanced measurement technology and microscale and nanoscale multiphase flow research in terms of modeling and algorithms. The third part reviews the current progress in multiphase flow-related research in nuclear engineering. Various aspects of solid-liquid, gas-liquid, and gas-particle flows, in particular, critical heat flux research and deep learning applications in nuclear engineering, have been extensively reviewed. This paper aims to explore multiphase flow research in the future in the nuclear and its related disciplines through the analysis and review of cutting-edge research and development states.
Article
Full-text available
We reveal mechanisms driving pre-filming wave formation of non-Newtonian banana puree inside a twin-fluid atomizer at a steam-puree mass ratio of 2.7%. Waves with a high blockage ratio form periodically at a frequency of 1000 Hz, where the collapse of one wave corresponds to the formation of another (i.e., no wave train). Wave formation and collapse occur at very regular intervals, while instabilities result in distinctly unique waves each cycle. The average wave angle and wavelength are 50{degree sign} and 0.7 nozzle diameters, respectively. Kelvin-Helmholtz instability (KHI) dominates during wave formation, while pressure effects dominate during wave collapse. Annular injection of the puree into the steam channel provides a wave pool, allowing KHI to deform the surface; then, steam shear and acceleration from decreased flow area lift the newly formed wave. The onset of flow separation appears to occur as the waves' rounded geometry transitions to a more pointed shape. Steam compression caused by wave sheltering increases pressure and temperature on the windward side of the wave, forcing both pressure and temperature to cycle with wave frequency. Wave growth peaks at the nozzle exit, at which point the pressure build-up overcomes inertia and surface tension to collapse and disintegrate the wave. Truncation of wave life by pressure build-up and shear-induced puree viscosity reduction is a prominent feature of the system, and steam turbulence does not contribute significantly to wave formation. The wave birth-death process creates bulk system pulsation, which in turn affects wave formation.
Article
Full-text available
Characterization of viscous, non-Newtonian atomization by means of internal waves is presented for a twin-fluid injector. Atomization of such fluids is challenging, especially at low gas-liquid mass ratios (GLRs). This paper details mechanisms that enhance their disintegration in a "wave-augmented atomization" process. The working fluid, banana puree, is shear-thinning and described by the Herschel-Bulkley model. Unlike a conventional airblast injector, an annular flow of banana puree is injected into a core steam flow, encouraging regular puree waves to form inside the nozzle. A pulsing flow develops with three distinct stages: stretch, bulge, and burst leading to an annular puree sheet stretching down from the nozzle exit. Rayleigh-Taylor instabilities and viscosity gradients destabilize the surface. During wave collapse, the puree sheet bulges radially outward and ruptures violently in a radial burst. Near-nozzle dynamics propagate axially as periodic fluctuations in Sauter mean diameter occur in a wave pattern. Numerical simulations reveal three atomization mechanisms that are a direct result of wave formation: 1) wave impact momentum, 2) pressure buildup, and 3) droplet breakaway. The first two are the forces that exploit puree sheet irregularities to drive rupture. The third occurs as rising waves penetrate the central steam flow; steam shear strips droplets off, and more droplets break away as the wave collapses and partially disintegrates. Waves collapse into the puree sheet with a radial momentum flux of 1.7 × 10 ⁵ kg/m-s ² , and wave-induced pressure buildup creates a large pressure gradient across the puree sheet prior to bursting.
Article
Full-text available
We present high-resolution three-dimensional (3-D) direct numerical simulations of breaking waves solving for the two-phase Navier–Stokes equations. We investigate the role of the Reynolds number ( Re , wave inertia relative to viscous effects) and Bond number ( Bo , wave scale over the capillary length) on the energy, bubble and droplet statistics of strong plunging breakers. We explore the asymptotic regimes at high Re and Bo , and compare with laboratory breaking waves. Energetically, the breaking wave transitions from laminar to 3-D turbulent flow on a time scale that depends on the turbulent Re up to a limiting value Reλ100Re_\lambda \sim 100 , consistent with the mixing transition in other canonical turbulent flows. We characterize the role of capillary effects on the impacting jet and ingested main cavity shape and subsequent fragmentation process, and extend the buoyant-energetic scaling from Deike et al. ( J. Fluid Mech. , vol. 801, 2016, pp. 91–129) to account for the cavity shape and its scale separation from the Hinze scale, rHr_H . We confirm two regimes in the bubble size distribution, N(r/rH)(r/rH)10/3N(r/r_H)\propto (r/r_H)^{-10/3} for r>rHr>r_H , and (r/rH)3/2\propto (r/r_H)^{-3/2} for r.Bubblesareresolveduptooneorderofmagnitudebelowr . Bubbles are resolved up to one order of magnitude below r_H , and we observe a good collapse of the numerical data compared to laboratory breaking waves (Deane & Stokes, Nature , vol. 418 (6900), 2002, pp. 839–844). We resolve droplet statistics at high Bo in good agreement with recent experiments (Erinin et al. , Geophys. Res. Lett. , vol. 46 (14), 2019, pp. 8244–8251), with a distribution shape close to N_d(r_d)\propto r_d^{-2}$ . The evolution of the droplet statistics appears controlled by the details of the impact process and subsequent splash-up. We discuss velocity distributions for the droplets, finding ejection velocities up to four times the phase speed of the wave, which are produced during the most intense splashing events of the breaking process.
Article
The focus of the present paper is to understand the large scale unsteadiness in the atomization of swirl coaxial jets and correlate it with the downstream spray statistics. A gas-centered swirl coaxial atomizer, in which a gaseous jet fragments an annular swirling liquid sheet, displays pulsating flow at certain momentum flux ratios (MFRs). These conditions were chosen to study the unsteady dynamics in the spray formation. Various zones of the spray such as near orifice region, primary atomization zone and far field spray were diagnosed using high-speed shadowgraphy technique. Proper orthogonal decomposition was employed on the time-resolved spray images to understand various unsteady modes. Three modes observed prominently were identified as large scale unsteady structures viz. axisymmetric pulsating mode, explosive mode and swirling mode. The pulsating mode was found to be more dominant in the pulsation regime, whereas the other modes gained significance at higher MFRs when the pulsation regime ceases to exist. Fourier analysis of temporal coefficients pertaining to pulsating mode showed a definitive frequency. The analysis of liquid shedding rate was found to be in synchronization with the pulsating mode which shows the upstream influence on periodic atomization. Spatio-temporal measurements of droplet sizes were carried out far from the atomizer. The temporal variation in the droplet number density and Sauter Mean Diameter depicted unsteady behavior; however, there is no preferred frequency in the power spectrum. This clearly indicates that the far field spray loses its memory of upstream periodic atomization in the pulsation regime of swirl coaxial atomizer.
Article
This study numerically investigated the atomization characteristics of misaligned impinging jets, with the misalignment ratio ê ranging between 0 and 0.2, by employing the volume of fluid method with an adaptive mesh refinement algorithm. The results show that the droplet Sauter mean diameter varies non-monotonically with ê and reaches the minimum value, which implies the best atomization performance, at ê=0.1 under operating conditions concerned in the present work. Meanwhile, the moderately misaligned impingement also leads to a more uniform spatial dispersion of the atomized fragments and droplets. These unique spray behaviors can be attributed to the instability and disintegration of the liquid sheet formed upon jet impingement, as evident from the non-monotonic dependence of the breakup length of the liquid sheet on the misalignment ratio ê. Analyses on the velocity fluctuation and vorticity distribution further suggest that the misalignment alters the intrinsic instability mode of the liquid sheet by introducing a lateral stretch effect, which diverts the peak streamwise momentum away from the centerline. The current finding indicates that misalignment tuning could be a promising optimization and control technique in propellant mixing and atomization.
Article
Studies on pressure-swirl atomizers have mainly focused on pressure-swirl atomizers with tangential input while there are limited studies on pressure-swirl atomizers with a spiral path. This study applies experimental and computational methods to provide a better understanding of flow development in this type of atomizer at the design point and outside the design point. Experimental results showed that as the pressure increases, the spray cone angle increases. This increase initially occurs with a higher slope and then the slope is toned down. While the drainage coefficient remains constant, the droplet diameter decreases as the pressure increases. It is observed that similar to the pressure-swirl atomizer with tangential input, the pressure-swirl atomizer with a spiral path has a conical hollow spray. At the constant mass flow rate, as the spiral path cross-section, the length of the swirl chamber and orifice diameter increase, the fluid film thickness and average diameter of droplets increase while the spray cone angle reduces. Further, increasing the number of spiral paths causes a wider spray cone angle, higher discharge coefficient, larger fluid film thickness, and larger droplet diameter. The results also showed that increasing the length of the orifice marginally affected the properties of the spray while significantly reducing the spray cone angle. It is important to note that the numerical results are in good agreement with the experimental data.
Article
Spray characterization in a novel twin-jet airblast injector is reported in this paper with the focus on the study of the effect of injector air swirl on droplet characteristics and dispersion behavior. The operational principle of the injector is based on achieving atomization of two liquid jets, injected in a radially opposite direction from a central hub by high-speed annular swirling cross-stream flow of air. Liquid jet atomization within model atomizers and the resulting spray study have not gained much attention in spite of its practical importance, for example, in lean premixed prevaporized combustors. In the present work, droplet size and three-component velocity measurements are measured in the above injector using the phase Doppler particle analyzer technique. Air velocity without liquid injection is also obtained using the laser Doppler velocimetry technique. For given inlet air and liquid mass flow rates, experiments are conducted in the absence and presence of annular air swirl corresponding to swirl number, S = 0 and 0.74, respectively. The addition of air swirl is found to dramatically affect the spray topology and also the measured spray characteristics as the droplet size reduces significantly downstream of the injector exit, which is explained. Droplet dispersion is studied by evaluating droplet size velocity correlation and also droplet Stokes number. The results not only provide insight into the physics behind improved atomization due to air swirl, but also demonstrate the ability of the novel injector to achieve atomization quality and high spray dispersion over a wide operating range.
Article
Two-fluid atomization of highly viscous and shear thinning non-Newtonian fluids must overcome the viscous resistance. Conventional air-blast atomizers address this requirement by operating at higher gas–liquid mass ratios (GLRs) at the cost of deteriorated efficiency. In this investigation, the atomization efficiency ηatom of a two-fluid atomizer is characterized. Fluidic design of the atomizer reduces the viscosity-induced dampening of disruptive inertia force and speeds up the breakup by providing kinetic energy of strategically positioned microjets of air. Three fluids of different viscosities and surface tensions were tested to examine the effects of properties on atomization efficiency. The presence of droplets and ligaments near the exit plane of the atomizer revealed that the atomizer enables prompt breakup of fluids without appreciable delay by viscosity. Consequently, it could break up complex non-Newtonian Jet A1 gel propellant with efficiency in the range of 0.015–0.1. Moreover, it is also capable of atomizing Newtonian fluids like water (high surface tension) with efficiency of 0.06–0.1 and Jet A1 (low surface tension) with efficiency of 0.02–0.08. Note that ηatom varied as GLRb (b=−0.73 to −0.77) for the liquids tested in this work. The results of this investigation show that the deterioration in the performance of internally impinging atomizers at higher air and liquid throughputs is not determined by the viscosity of liquid but is predominantly due to the loss in kinetic energy of the excess air.