Sei sulla pagina 1di 18

Received: 28 November 2018 Revised: 20 March 2019 Accepted: 11 May 2019

DOI: 10.1002/aic.16675

TRANSPORT PHENOMENA AND FLUID MECHANICS

Flow characteristics of vortex based cavitation devices


Computational investigation on influence of operating parameters and scale

Alister Simpson | Vivek V. Ranade

School of Chemistry and Chemical


Engineering, Queen's University Belfast, Abstract
Belfast, UK Vortex based hydrodynamic cavitation (HC) devices offer various advantages over
Correspondence conventional linear flow devices, such as early inception, low erosion risk, and higher
Vivek V. Ranade, School of Chemistry and cavitational yield. Despite several promising applications, the key underlying flow
Chemical Engineering, Queen's University
Belfast, Belfast BT9 5AG, UK. characteristics are not yet adequately understood. This article presents results of a
Email: v.ranade@qub.ac.uk computational investigation into cavitation in vortex devices. Multiphase computa-
tional fluid dynamics results are presented and compared with experimental data on
pressure drop over a range of flow rates. The results highlight the unique hydrody-
namic characteristics of this type of device in relation to conventional linear flow
reactors; cavitation inception occurs in the liquid bulk away from sold surfaces, and
rapid pressure recovery rates are achieved. The models were used to simulate
detailed time–pressure histories for individual vapor cavities, including turbulent fluc-
tuations. The developed approach, models and results will provide a sound and useful
basis for comprehensive multiscale modeling of vortex-based devices for HC.

KEYWORDS
CFD, multiphase, scale up, turbulence models, vortex cavitation

1 | I N T RO D UC T I O N associated radial reduction in static pressure. With sufficiently high


tangential velocities, the local pressure at the core of the vortex
Vortex based flow devices have found extensive use in a wide variety of reaches the vapor pressure of the working fluid. Vapor cavities
1,2
industrial applications, most notably in cyclonic separation systems. formed at these low pressures may experience oscillatory growth,
Recent novel examples of vortex flow devices in chemical reaction engi- eventually collapsing to generate spots of extreme temperatures and
3-5
neering include rotating fluidized bed reactors, in which particle bed flu- pressures (5,000 K and 1,000 atm, respectively6,7). These collapse
idization is achieved by generating highly swirling flow within a fixed events are accompanied by localized shock waves and extremely high
geometry reactor. Another promising industrial application for vortex- shear rates, and as a result cavitation can be famously damaging in
based reactors is in hydrodynamic cavitation (HC) applications, where the pressurized fluid handling equipment. When carefully harnessed, how-
intense collapse of vapor cavities in a flowing liquid medium can be utilized ever, these effects can be utilized to produce intense reactions; at the
to facilitate a range of industrial processes. The focus of the present study point of collapse, the contents of the cavity dissociate and release
is a novel type of vortex cavitation reactor, based on a vortex diode. Vor- highly oxidizing radical species, including hydroxyl radicals (OH), which
tex diodes are bidirectional fluid flow devices, characterized by a high flow can be utilized in advanced oxidation processes.
resistance in one direction relative to the other. The basic construction The successful use of HC has been reported for a wide range of
features a tangential inlet, swirl chamber and an axial outlet port (Figure 1). industrial applications; for wastewater in particular it has been identi-
In “reverse” operation, flow enters through the tangential inlet, fied as a promising method for the removal of micropollutants, and
which quickly establishes a vortex in the main swirl chamber with an numerous studies have highlighted its potential to degrade a range of

AIChE Journal. 2019;e16675. wileyonlinelibrary.com/journal/aic © 2019 American Institute of Chemical Engineers 1 of 18


https://doi.org/10.1002/aic.16675
2 of 18 SIMPSON AND RANADE

complexities in vortex diodes. They described the formation of an


inherently unsteady, precessing vortex core (PVC) along the device
axis, similar to behavior widely reported in vortex separation units.2,25
Pandare and Ranade also reported a significant core of reverse flow in
the axial port, which persists for up to 30 diameters into the down-
stream pipework. Similar observations of reverse flow in vortex units
have been reported by Niyogi et al,5 who presented a detailed
description of secondary flow patterns in single phase gas vortex
devices of similar geometry.
Vortex cavitation itself has been the subject of numerous studies,
primarily in the context of tip vortices formed at the trailing edge of
hydrofoils. This specific form of vortex formation is understood to be
a major contributor to erosion in hydraulic machinery, as well as noise
and vibration. Arndt et al have highlighted some of the complexities
of such cavitating vortices, which exhibit a range of broadband pres-
F I G U R E 1 Vortex diode schematic [Color figure can be viewed at sure fluctuations and complex collapse characteristics.26,27 Recent
wileyonlinelibrary.com] computational modeling efforts have highlighted novel ways to pre-
dict the formation and evolution of cavitating line vortices, such as
contaminants recalcitrant to conventional treatment methods. Com- the combined lagrangian-vof approach developed by Ma et al,28 and
prehensive summaries can be found in the reviews compiled by more recently a modeling approach has been presented by Chen et al,
8 9
Rajoriya et al, Dular et al, and Ranade and Bhandari. 10
To highlight which offers a method to predict tip vortex cavitation inception by
some specific examples, Kalumuck and Chahine 11
conducted experi- accounting for water quality.29 Confined cavitating vortex flows, as
ments to degrade p-nitrophenol (pNP) using a recirculating loop cavi- found in vortex diodes, have received relatively little attention on the
tation system, and reported reductions in concentration of the order other hand, and advances in device design depend on developing an
of 90%. Importantly, the energy efficiency was found to be two orders improved understanding of the multiphase flow behavior. The present
of magnitude greater than ultrasonic cavitation. Capocelli et al 12
have work aims to address these gaps in current understanding, using a
reported similar success in degrading pNP using HC, with removal series of numerical investigations to highlight the unique hydrody-
rates of over 40% achieved using a venturi loop reactor. Vortex cavi- namic characteristics of the device. We present a detailed description
tation reactors have recently been reported to yield promising results of multiphase flow with phase change for cavitating vortex devices
in the degradation of industrial effluents; Suryawanshi et al 13
have across a range of operating conditions, and how these characteristics
demonstrated the effectiveness of vortex diode reactors in degrading vary with device scale up.
industrial solvents, achieving 80% toluene removal. Beyond wastewa-
ter applications, the successful adoption of HC technology has been 2 | M A T H E M A T I C A L M OD E L S
reported in a wide range of industrial processes, including microbial
disinfection,14 fuel desulfurization,15 biodiesel synthesis,16,17 biomass 2.1 | Flow and turbulence
pretreatment,18,19 and in food and beverage production20,21 to name
just a few. The working medium is treated as a single fluid, comprised of a homo-

The published literature on HC applications features a range of geneous mixture of two phases. Throughout the present study a

nonoptimized reactor designs, most typically of venturi or orifice type Reynolds averaged Navier Stokes (RANS) approach was adopted to

construction. Cavitation in such devices initiates and evolves around model turbulence effects. In this approach the velocity terms in the

the surface of the orifice or venturi restriction itself,22,23 which governing equations are replaced by the sum of their mean (uÞ and
0

imposes some inherent limitations; first, the reactor surfaces are instantaneous (u ) components, u = u + u0 , and an ensemble average
exposed to potential cavitation erosion, which can greatly reduce the taken. The additional terms that result are known as the Reynolds
 
operational lifetime and performance of the device. Second, the stresses, which have the general form ∂xj − ρui uj . In order to close
restriction itself creates a clogging risk for applications involving solids the momentum equation, additional physical models are required to
loading, such as biomass pretreatment. Vortex devices can potentially approximate these additional Reynolds stresses. One approach is to
overcome these limitations, as cavitation occurs in the core of the use a Reynolds stress model (RSM), which involves solving separate
fluid bulk as opposed to the surface of the restriction. With the cor- transport equations for each of the additional Reynolds stresses (six in
rect design approach there is the potential to harness the benefits of total for 3D cases). Full details of this approach can be found in Wil-
cavitation while eliminating the risks of clogging and erosion. The flow cox.30 More typical in RANS approaches is to employ the Boussinesq
24
in such devices is however highly complex; Pandare and Ranade hypothesis, which approximately relates the Reynolds stresses to the
presented a detailed description of the single phase flow features and mean velocity gradients in the flow as follows:
SIMPSON AND RANADE 3 of 18

 
∂ui ∂uj 2 mass fraction, f, computed locally using a separate transport
−ρui uj = μt + − ρkδij : ð1Þ
∂j ∂i 3 equation:

∂  ! 
This expression introduces the scalar quantity μt, the turbulent viscos- ðρm f Þ + r  ρm v m f = r  ðΓrf Þ + Re −Rc : ð7Þ
∂t
ity. This quantity is typically modeled through additional transport
equations for the turbulent kinetic energy (k) and turbulent dissipation Here ρm, ρv, and ρl refer to the densities of the mixture, vapor and liq-
rate (ε), or the specific dissipation rate (ω). In the present work, the uid, respectively. An additional pair of mass source and sink terms is
shear stress transport (SST) k-ω model of Menter31 has been selected now introduced for the evaporation (Re) and condensation (Rc) of the
for use, in which the turbulent viscosity, μt is defined as: vapor phase. Numerous modeled source and sink terms have been
proposed to account for cavitation mass transfer in Equation (7), with
ρk the majority of approaches deriving these terms from a reduced form
μt = α , ð2Þ
ω
of the Rayleigh–Plesset equation, commonly shortened to the “R-P”
where the premultiplier α is typically computed from a number of further equation:
subfunctions and empirical constants, presented in detail in References
31
and 32. The turbulence kinetic energy and specific dissipation rate are "   #
d2 RB 3 dRB 2 2σ 4μ dRB
then calculated by two separate transport equations: ρ RB + = pB − p− − : ð8Þ
dt2 2 dt RB RB dt

 
∂ ∂ ∂ ∂k
ðρkÞ + ðρkui Þ = Γk + Gk − Y k + Sk , ð3Þ In this expression, RB is the bubble radius, pB refers to bubble surface
∂t ∂xi ∂xj ∂xj
pressure, and p is the local liquid phase pressure. Full derivation of the
  R-P equation is presented in Reference 35
, and numerous examples of
∂ ∂ ∂ ∂ω
ðρωÞ + ðρωui Þ = Γω + Gω − Y ω + Sω , ð4Þ solutions of the R-P equation exist in literature, for example Alehossein
∂t ∂xi ∂xj ∂xj
and Qin.36 Commonly, when modeling cavitation mass transfer mecha-
nisms in computational fluid dynamics (CFDs) codes, the surface ten-
G represents the generation, Γ the effective diffusivity, Y is the dissi-
33 sion, viscous damping and higher order acceleration terms in
pation due to turbulence, and S is a user defined source term. The
Equation (8) are neglected to produce a mass transfer rate term of the
subscripts k and ω denote turbulence production and dissipation,
following general form:
respectively. Standard two-equation turbulence models are known to
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
suffer from excessive predictions of turbulence kinetic energy in the
dRB 2 ðpB − pÞ
vicinity of stagnation points, which is particularly relevant to swirling ffi ð − 1Þn : ð9Þ
dt 3 ρl
flows as the swirl-axis itself represents a stagnation condition. Limit-
ing formulations for the production term, Gk, have therefore been pro-
There exists in open literature a variety of source term approximations
posed by Menter,32 and also Kato and Launder.34 The formulation
of this general form, for example, those proposed by Schnerr and
proposed by Menter is as follows:
Sauer,37 Zwart et al,38 and Singhal et al.39 The Singhal model was cho-
sen for use in the present study, given below in Equations (10) and (11).
Gk = min ½Gk , Clim ρk, ð5Þ
pffiffiffi  

k 2 pv − p 1=2
where the limiting coefficient, Clim is set as 0.9 in the original publication. Re = C1 ρv ρl ð1 − f v −f g Þ, ð10Þ
σ 3 ρl
Kato and Launder proposed an additional limiting correction on the pro-
duction term, based on the observation that high turbulence kinetic pffiffiffi  

k 2 p− pv 1=2
energy production in stagnation regions is generally caused by the calcu- Rc = C2 ρv ρl fv: ð11Þ
σ 3 ρl
lation of excessive shear strain rate, γ. As such, this is corrected by the
local vorticity rate, Ω, which tends to low values in stagnation regions:
Real engineering liquid systems typically contain a small quantity of
noncondensable gas (NCG), present in a dissolved state and poten-
Gk = μt γΩ: ð6Þ
tially as bubbles introduced by aeration, which can influence the
observed cavitation behavior.40,41 In Equation 10, NCG is accounted
The sensitivity of the results to the choice of closure model is
for through the additional mass rate term, fg. The precise amount of
presented in full detail in the subsequent turbulence model sensitivity
NCG is typically an unknown in most engineering applications how-
section. ever, as direct measurements are extremely difficult. Solubility data
for oxygen in water gives an equilibrium concentration of 8 g/m3
(8 ppm) at typical laboratory room conditions of 1 atm and 22 C,
2.2 | Cavitation model
however the total gas content can vary according to a wide range of
The two-phase flow field in this work is modeled by assuming a additional factors, such as prior liquid processing and storage steps,
homogeneous mixture of water liquid and vapor, with the vapor for example; degassing, pressurization, and the length of time the
4 of 18 SIMPSON AND RANADE

system has been exposed to air. Liquid vortex units in particular are recommended values have been adopted throughout the present
known to feature a distinct core of NCG above the vapor limit of the study without modification.
working fluid.2,25,29,42,43 In the present investigation it is assumed that
the liquid system free surface is exposed to atmosphere for a suffi-
3 | RESULTS AND DISCUSSION
cient length of time for it to reach equilibrium with air at atmospheric
pressure and room temperature, and that the system has also had suf-
3.1 | Diode configurations
ficient time to settle such that any entrained larger macroscopic bub-
bles rise to the free surface and escape. As such, the fixed fraction of Figure 2 presents the construction and dimensions of the vortex
NCG has been set at equilibrium saturation conditions for air in water devices considered in the present study. Four geometrically similar
throughout. units were studied of different scales, the smallest having a 6 mm
The evaporation and condensation source terms in Equa- throat diameter (dt), and largest having a 48 mm throat diameter.
tions (10) and (11) also feature a pair of premultipliers, C1 and C2.
Empirical premultipliers are also incorporated into other cavitation
3.2 | Numerics and convergence strategy
models of this general type based on this general “reduced Rayleigh
Plesset” approach described by Equation (9). The necessity of empiri- The model equations described in the preceding section were all solved
cal constants of this type suggests, first, that the bubble growth rate using commercial CFD code, Ansys Fluent (v17). Following similar
term does not directly correlate with the magnitudes of cavitation approaches adopted by Niyogi et al,5 the present study primarily makes
mass transfer rates observed in reality, and also, second, that the basic use of a 2D axisymmetric approach. This simplification necessarily mis-
approach implies equivalent evaporation and condensation rates, ses some of the detail revealed by a time-resolved 3D calculation, and
which requires additional empirical correction. In validating their cavi- with this in mind an initial series of transient 3D CFD calculations were
tation calculations, Singhal et al39 compared predictions against avail- performed in order to derive appropriate boundary conditions for the
able published empirical data sets for a range of orifice, hydrofoil, 2D models. This reduced 2D approach subsequently allowed a larger
blunt body and impeller flows, and stated that hundreds of permuta- number of operating conditions and parameters to be explored within
tions were considered in order to determine suitable values for this feasible computational timescales. A detailed comparison of the 2D and
pair of constants in order to reliably and repeatedly match experimen- 3D approaches is presented and discussed in the subsequent sections.
tal values of flow rates, discharge coefficients and cavitating flow pat- Although the 2D simulation results presented here do not exhibit promi-
terns. The values of the empirical constants C1 and C2 were set as nent unsteady behavior in the predicted velocity and pressure fields, an
0.02 and 0.01, respectively, in the original publication, and these unsteady solver was required in order to ensure convergence across all

FIGURE 2 Vortex device geometry and operating conditions [Color figure can be viewed at wileyonlinelibrary.com]
SIMPSON AND RANADE 5 of 18

of the cases studied. Time steps of the order of 1e−4 to 1e−6 were used swirl velocity and turbulent kinetic energy through the midsection of
across the range of operating conditions set out in Figure 2. The SIMPLE the smaller device chamber (Device A) are compared for four different
algorithm was used for pressure velocity coupling, with the PRESTO! mesh sizes in Figure 3b and c, respectively. Negligible differences
Discretization scheme applied for pressure. For the momentum and tur- were found in the predictions above an overall mesh count of
bulent quantities second-order discretization was applied in each 120,000 cells, which was therefore set as the grid count for the
instance. remainder of the present study for Device A. This featured a cell size
of 0.08 mm (= dt/75) through the center of the device and axial port,
and a near wall initial cell height of 0.003 mm, ensuring y+ values of
3.3 | Mesh and boundary conditions
unity even at the maximum simulated flow rate. For the larger
At the outset of the 2D simulations, a comprehensive grid sensitivity geometries B, C and D, the near wall sizing was held constant, and the
study was performed to establish a suitable level of mesh refinement mesh spacing adjusted so as to result in final mesh sizes which scaled
in the regions of particular interest; Figure 3 illustrates the mesh approximately with device scale. The highest cell count for the largest
topology and refinement levels. The boundary conditions are also 48 mm device was 840,000 elements.
shown; at the inlet, velocity profile data extracted from 3D simulation
results was applied. The outlet of the device in practice discharges
3.4 | Turbulence closure model sensitivity
freely to a holding tank, and this was represented by setting a fixed
pressure equal to 1 atm at the outlet boundary, positioned 40× diam- The interaction between cavitation and turbulent flow structures is
eters downstream of the axial port exit. The computed gradients of extremely complex, and the detailed mechanisms are not yet fully

F I G U R E 3 (a) Computational grid and boundary condition recipe. (b) Mesh sensitivity study; swirl velocity. (c) Mesh sensitivity study;
turbulent kinetic energy [Color figure can be viewed at wileyonlinelibrary.com]
6 of 18 SIMPSON AND RANADE

understood. The difficulties in turbulence modeling in cavitating flows combined turbulence production limiting functions of Menter, and
and proposed approaches have been well documented in open litera- Kato and Launder presented previously in Equations (5) and (6).
ture44-47; for detached flows in particular, such as those produced over Comparisons of the swirl velocity profiles using each approach are
hydrofoils, there is strong interaction between the eddy vortices presented in Figure 4. The results indicate that the k-ε model under-
formed and the generated vapor phase, creating strong pressure fluctu- predicts the swirl magnitudes in the current case, owing to an over-
ations.48-50 These complex turbulent features present significant prediction of turbulence kinetic energy. This result is somewhat
modeling challenges; In studying a cavitating venturi, Reboud et al44 expected owing to the widely reported deficiencies of ε-based models
made the argument that the standard k-ε model tends to produce high in similar flow scenarios57,58; perhaps less expected however is the
turbulent viscosity (μt) predictions in the separation region, which acts difference in the unmodified SST k-ω model predictions in 2D and 3D.
to dampen unsteady effects, leading to the prediction of a stable, fixed The 2D SST k-ω models show significantly lower tangential velocities,
separation bubble. To properly capture unsteady cavity shedding, the which correspondingly mean that the minimum pressure, and thus
authors found it necessary to apply an essentially arbitrary premultiplier cavitation inception, at the axis of the swirl chamber is similarly
to limit the turbulent viscosity. A number of examples exist in open lit- under-predicted. This suggests that the imposition of zero velocities
erature of favorable correlations being obtained between models apply- in the radial and swirl directions at the axis leads to an overprediction
ing this type of empirical limiter with experimental data for unsteady of turbulence production in 2D, which necessitates the introduction
cavitation.45,51-54 For detached venturi flows Charriere et al46,47 have of additional production limiting terms. Imposing the combined limit-
successfully demonstrated that the use of a Reboud type eddy- ing formulations presented in Equations (5) and (6) to the production
viscosity limiter in RANS models can reliably predict the re-entrant jet term in the SST k-ω model was found to be successful in achieving
phenomenon for 2D venturi configurations.46 The arbitrary nature of much closer agreement, producing results in line with the RSM model.
the applied correction factors however means that it is likely to be This approach was then subsequently adopted across the full range of
dependent on geometry, and in lieu of sufficient experimental data to operating conditions, and produced good agreement with experimen-
“tune” the turbulence model for the present flow case no such addi- tal data across the investigated range, and at different scales (see
tional limiters have been applied directly to the turbulent viscosity term experimental comparison shown in Figure 7 in the following section).
in this study. More recent studies on hydrofoils highlight the highly
three-dimensional, complex interactions between detached eddy vorti-
ces and the secondary gas phase, and in such cases some degree of
3.5 | Comparison of 2D and 3D simulations
scale resolving is required to obtain reliable results.55,56 To judge the capability of the 2D approach to capture the key hydro-
In the case of a confined, single vortex of the type studied in the dynamic characteristics of the device, an initial set of 3D, time-
present work cavitation is initiated and develops along the swirl axis, resolved CFD calculations were performed. The 3D calculations were
and the interactions between turbulence and the secondary phase in performed on a hexahedral grid of 4.1 million elements, which
this scenario are not well understood. At the axis of the vortex chamber resulted from a separate mesh independency study. The 3D results
there is no separation, or eddy vortex shedding, however a stagnation were initially used to derive appropriate boundary conditions for the
region of a different nature is present; tangential velocities go to zero 2D models. Velocities at the inlet plane of the 2D models were sam-
at the axis, and the axial velocities also tend to low/near zero values. As pled from the time averaged 3D results at radius r = ri, (Figure 5), and
such, as is the case for cases with detached eddy shedding, these stag- nondimensional radial and tangential velocity profiles then extracted
nation regions require careful consideration when imposing a turbu-
lence closure model so as to avoid overprediction of the production
term. Additionally, a recent study by Zaman et al57 has highlighted the
fact that constrained highly swirling flows can lead to re-laminarization
at nominally turbulent Reynolds numbers; as a consequence it could be
anticipated that standard two-equation approaches such as the k-ε
model would also give rise to significant discrepancies in single vortex
flows. In the present study it is particularly crucial that the choice of
turbulence closure model leads to realistic tangential velocity profiles,
as this tangential acceleration produces the low-pressure regions along
the axis of the device that drive cavitation mass transfer.
In the present case, a number of different modeling approaches
were evaluated and compared with full 3D predictions. The models
evaluated include an ω-based RSM,30 a laminar calculation, the
Renormalization group (RNG) k-ε as well as two different variations of
F I G U R E 4 Swirl velocity profiles in chamber midplane—
the SST k-ω model. The two SST k-ω variants comprise the turbulence model comparison [Color figure can be viewed at
unmodified SST k-ω, and an SST k-ω model incorporating the wileyonlinelibrary.com]
SIMPSON AND RANADE 7 of 18

to be applied as boundary conditions for 2D axisymmetric models. the multiphase predictions. This is in contrast to linear flow devices
These inlet profiles were then scaled accordingly to simulate at a such as orifice and venturi, where the onset of cavitation is accompa-
range of operating conditions. One prominent feature of the flow con- nied by a reduction in flow rate at a given pressure ratio in compari-
ditions illustrated in Figure 5 is the near-zero radially inward mass son to single phase predictions.22,61,62 In such linear flow devices,
flow across a significant portion of the span at the sample radius/2D cavitation initiates and grows at the restriction. The gas phase there-
inlet radius, ri, with the majority of the mass flux concentrated toward fore presents an effective reduction in liquid flow area, which acts to
the outer walls of the swirl chamber. reduce the flow rate in the cavitating regime. In a vortex diode, cavita-
The predicted swirl strength through the midplane of the swirl tion initiates along the axis of the chamber, rather than at the device
chamber is compared for the 3D and 2D approaches in Figure 6. One throat, and as the results in Figure 6 imply the gas phase therefore
notable omission from the 2D predictions is vortex precession, which
is a common feature of vortex flows.59,60 Previous studies of single-
phase flow in vortex diodes has reported the presence of a PVC, with
a frequency of the order of 60 Hz.24 The 3D results in Figure 6 are
assembled from a time average, and despite ignoring precession
effects close alignment is found with the 2D axisymmetric approach
in both the maximum swirl ratio and gradient toward the center of the
device. Figure 6b presents a similar comparison of the corresponding
predictions of static pressure through the chamber midplane, again
showing close agreement between the two modeling approaches. The
comparison plots highlight a slight eccentricity of the minimum pres-
sure core found by the 3D approach; the tangential inlet in this
instance generates a circumferential bias in mass flux, which pushes
the low-pressure core slightly off-axis. Despite ignoring these circum-
ferential nonuniformities, the 2D approach shows good general agree-
ment between the axisymmetric model and the full 3D CFD approach
in terms of the magnitude of pressure reduction at the vortex core,
which is crucial in the study of cavitation inception and evolution
within the vortex device.

3.6 | Comparison with experimental data


Figure 7 presents the predicted pressure drop, Δp, across Devices A
and B with increasing throat velocity, ut, alongside experimentally
measured values. The 2D predictions provide close agreement at both
F I G U R E 6 (a) Predicted swirl ratio in chamber midplane; 2D
device scales. In addition to results from 2D multiphase simulations, versus 3D model. (b) Predicted static pressure in chamber midplane;
results from corresponding single phase calculations are also included 2D versus 3D model [Color figure can be viewed at
in Figure 7 for Device A, showing the characteristic curve to overlap wileyonlinelibrary.com]

F I G U R E 5 Two-dimensional
approximation of vortex unit: velocity
profiles at the inlet [Color figure can
be viewed at wileyonlinelibrary.com]
8 of 18 SIMPSON AND RANADE

visualization and particle image velocimetry measurements of cavitating


flows in vortex devices. In order to inform the design of an appropriate
experimental facility and measurement strategy, it is first important to
understand the key flow features, and the subsequent sections therefore
describe numerical studies which span a wide parameter space (geomet-
ric as well as operational) to elucidate these key flow features. These
results are also qualitatively compared with published studies on the key
flow features of geometrically similar gas vortex units,5 and any observed
similarities and differences highlighted.

3.7 | Predicted flow fields


Figures 9a and 9b presents in plane streamlines and velocity vectors
F I G U R E 7 Measured and predicted Δp with increasing throat to illustrate the flow through the device chamber and axial port at
velocity, ut [Color figure can be viewed at wileyonlinelibrary.com] two conditions: a throat velocity, ut of 2.3 m/s, and at ut = 3.5 m/s,
which is well into the cavitating regime. The velocity vectors have
been normalized to better depict the local flow directions; it should
be noted that this results in small magnitudes of reverse flow near
the inlet plane being represented by vectors which project out of
the domain; this exaggerates the fact that radial velocities are near
zero across most of the inlet span, with locally high inward velocities
along the outer chamber walls. Figure 9c presents results for the
larger scale Device B at a throat velocity of 3.5 m/s. The presence
of two counter-rotating secondary flow vortices is shown in the
chamber, generated by the high inward radial velocities along the
chamber walls. This observed secondary flow pattern is similar to
that found for gas vortex units by Niyogi et al5 Also highlighted by
F I G U R E 8 Predicted cavitation inception versus throat velocity,
Niyogi et al was the formation of a backflow region in the axial port,
ut [Color figure can be viewed at wileyonlinelibrary.com]
also indicated in Figure 9, with pockets of negative axial velocity
predicted along the axis just after the diverging section, persisting
does not limit the flow rate over the range of conditions studied here. up to 7 diameters downstream. The geometry studied by Niyogi
As detailed in the subsequent results section, predictions indicate that et al featured a constant cross section axial port, and as such the
the cavitating region does not extend to fill the axial port even at the backflow region was found to extend from the chamber back wall
highest flow rates studied here, and it may be possible that higher through to the observed exit; in the present case the variation in
flow rates could lead to a deviation in this trend. For the present case, cross sectional area in the axial port is predicted to interrupt the for-
cavitation inception is indicated in Figure 8, which shows both the mation of a backflow region. Comparing Figures 9a and 9b, the flow
minimum pressure (dotted line) and cavitation number, σ (solid line),
pattern predictions show little change in general behavior with
as a function of throat velocity. In this case the cavitation number is
increasing throat velocity. The nondimensional numerical values
defined by Equation (12), with the maximum tangential velocity
remain consistent as throat velocity increases, reflecting a linear
applied as the characteristic velocity. Using this definition, cavitation
increase in the absolute swirl velocity with increasing throat veloc-
numbers of 1 were found to correlate to the minimum pressure in the
ity. Similarly, the absolute strength of the reverse flow core
device reaching the vapor pressure, pv. Cavitation inception points
increases linearly with increasing throat velocity. The non-
at both scales are similar, with inception for the larger Device B
dimensional values shown in Figure 9c at the larger scale factor of
predicted slightly earlier at throat velocities of 2.5 m/s, whereas
2 are also consistent with the smaller device scale, suggesting that
inception is indicated at 2.7 m/s for Device A.
the general flow structure, in terms of the strength of the swirl com-
p −p
σ=1 2 2 v : ð12Þ ponent and reverse flow core, remains a linear function of the char-
2 ρuθ max acteristic throat velocity as the device scale increases.
In addition to comparing computed and measured flow rate and pressure The normalized tangential velocity profiles through the midplane of
losses, ideally it would be useful to compare predicted velocity and tur- the chamber are presented in Figure 10 at each device scale. The predic-
bulent kinetic energy profiles, however no such experimental data cur- tions show the smallest Device A to exhibit the lowest maximum tangen-
rently exists in open literature for multiphase vortex units. Further tial velocities, with a relatively large increase in maximum swirl ratio
experimental investigations are planned to carry out detailed flow predicted with a doubling of device scale to Device B. As further
SIMPSON AND RANADE 9 of 18

F I G U R E 9 (a) Device A |
ut = 2.3 m/s| Re = 13,775. (b) Device A
|ut = 3.5 m/s| Re = 20,960. (c) Device B
|ut = 3.5 m/s| Re = 41,925 [Color figure
can be viewed at
wileyonlinelibrary.com]

highlighted in Figure 10b, the maximum values are then indicated to pla- increases, the minimum pressure is predicted to reach the saturated
teau with increasing scale beyond Device C (dt = 24, scaling factor = 4). vapor pressure at a throat velocity of 2.5 m/s. This identifies the cavi-
tation inception point, beyond which the low-pressure vapor core
continues to extend along the axis with increasing flow rate. Above
3.8 | Device pressure recovery
throat velocities of 3 m/s, the extent of the low-pressure core is
The evolution of the absolute static pressure along the axis of the found to be limited by the increase in cross section, and the majority
small vortex Device A is illustrated in Figure 11. As throat velocity of the pressure recovery in the axial direction is found to occur over a
10 of 18 SIMPSON AND RANADE

relatively short distance either side of this expansion point (between


z = 3× chamber height, H and z = 4× H). The influence of device scale
is also included in Figure 11b, with the predictions showing a marginal
increase in the axial extent of the low-pressure region for Device B
(2× scale); this can be attributed to a rise in overall pressure loss with
increasing device scale, as determined experimentally between
Devices A and B (Figure 7).
Unlike linear flow devices, such as venturi and orifice type cavi-
tation devices, the axial pressure recovery is not the dominant pres-
sure gradient; Figure 12a presents the radial pressure distribution
at a series of axial locations through the device, with the
corresponding axial pressure gradient provided in Figure 12b for
comparison. The radial pressure gradient is found to remain rela-
tively constant up until an axial distance equal to 3× H, again just
prior to the increase in cross section. Comparing this to the maxi-
mum axial pressure gradient shows the radial gradient to be higher
by a factor of 3. Smaller cavities will tend to move between the
vapor core to the outer wall, and may therefore experience a sharp
rise in static pressure in the axial port of up to 30× pv, which equa-
tes to double the outlet/fully recovered pressure. Contrastingly, in
linear flow devices, the maximum pressure gradient experienced by
individual cavities is limited to the difference in vapor pressure and
the device outlet pressure.
Research in this field has found that increasing the rate of pres-
F I G U R E 1 0 (a) Chamber swirl velocity profiles at different device sure recovery has a positive effect on device performance in terms
scale. (b) Maximum swirl velocities versus Reynolds number at of increasing the intensity of cavitation bubble collapse. Particularly
different device scales [Color figure can be viewed at
relevant is a study by Soyama et al,63 who found that increasing the
wileyonlinelibrary.com]
back pressure in a venturi cavitation reactor by a factor of 2 resulted

F I G U R E 1 1 (a) Static pressure contours,


Diode A. (b) Static pressure distributions
along device axis [Color figure can be
viewed at wileyonlinelibrary.com]
SIMPSON AND RANADE 11 of 18

(a)

(b)

FIGURE 12 Predicted static pressure distributions (Diode B, ut = 3.3 m/s) [Color figure can be viewed at wileyonlinelibrary.com]

in a recorded increase in acoustic power of two orders of magnitude. and the increase in cross section area, owing to the rapid recovery
Capocelli et al64 have also studied the effect of increasing back pres- in pressure at this location. Figure 13b shows the gas holdup, εg, as
sure in a venturi cavitation reactor on the degradation of pNP in a function of throat velocity at each device scale, where εg is the
water. For the same overall device pressure drop, an increase in back ratio of gas volume to total reactor volume. The predictions indicate
pressure of the order of 0.4 bar was reported to deliver an apprecia- a plateau beyond throat velocities of 4.0 m/s, which corresponds to
ble improvement in degradation performance, with recorded con- the cavitating vapor core reaching the area change in the axial port.
centration ratio found to reduce by 15%. The rapid pressure With increasing size, corresponding to the slight increase in extent
recovery attained in the vortex unit axial port will therefore in princi- of the low-pressure region, there is an accompanying marginal
ple contribute to achieving more intense cavity collapse, and by uti- increase in gas holdup. In general, the results suggest that the cavi-
lizing the radial pressure gradient the performance of the vortex tation generation rate is a simple function of device flow rate, which
device may be optimized for practical applications. The effect of in this reactor geometry is predicted to reach an asymptote at
these predicted pressure gradients on the pressure histories experi- throat velocities around 4 m/s.
enced by individual cavities is explored further in Section 3.11.

3.10 | Predicted turbulent fluctuations


3.9 | Phase change predictions
Individual cavities experience oscillatory expansion and compression
The corresponding predicted vapor volumes for Device A are given as a result of the turbulent fluctuations in pressure experienced over
for two different flow rates in Figure 13a, showing the evolution of their lifetimes. This turbulent pressure history dictates the final peak
the gas and vapor in the vortex core of the unit. The key feature of temperatures and pressures at the point of collapse; as such it is vital
the two-phase flow in the vortex device is the cavitating vapor core, to understand the magnitudes of the turbulence properties in and
which is constrained to the center of the device, away from the around the cavitating vortex core region. The turbulence kinetic
outer walls of the axial port. The extent of the vapor core is energy in Device A is given in Figure 14a, normalized by the square of
restricted to the portion of the axial chamber between the rear wall throat velocity, with the corresponding normalized turbulence eddy
12 of 18 SIMPSON AND RANADE

frequencies plotted in Figure 14b. Turbulence kinetic energy magni- throat velocity is increased, the turbulent kinetic energy magnitudes
tudes are highest in the axial port at the point of increase in cross sec- show a proportional increase as expected, with the highest magni-
tional area. At this location there are high gradients of both swirl tudes observed at the change in cross section in the axial port.
velocity and axial velocity, with a strong reverse flow along the central Figure 14b shows the corresponding turbulence eddy frequency dis-
axis, all of which combine to create a highly turbulent region. As tributions, which indicate high values at the rear wall of the swirl

F I G U R E 1 3 Predicted gas
holdup [Color figure can be
viewed at wileyonlinelibrary.com]

F I G U R E 1 4 (a) Turbulent kinetic energy, Device A. (b) Turbulence eddy frequencies, Device A [Color figure can be viewed at
wileyonlinelibrary.com]
SIMPSON AND RANADE 13 of 18

chamber; for Device A at a throat velocity of 3 m/s, the absolute more detailed discrete phase calculations; which are discussed in
values at this location are of the order of 500 kHz. Immediately down- detail in the following section.
stream of the change in cross section in the axial port, the absolute
values of eddy frequency correspond to 5 kHz.
3.11 | Discrete cavity trajectories
The influence of device scale on turbulence properties is pres-
ented in Figure 15. Some scale effects are immediately evident; tur- The overall reactor performance is governed by complex interactions
bulent kinetic energy increases by the order of 40% with increasing of numerous phenomena, including the cavity generation rate, cavity
size between the smallest and largest geometries studied. The size distributions, the turbulent pressures experience by the cavities,
trend in eddy frequency magnitudes with on the other hand shows as well as cavity–cavity interactions. The predicted flow field results
an inversely proportional relationship to device scale. At the largest therefore form just one part of a very complex picture. Final cavity
scale with throat diameter, dt, of 48 mm, the absolute values of collapse conditions are currently determined via direct numerical sim-
12
eddy frequencies in the axial port reduce to 0.6 kHz. In terms of ulation of the Rayleigh–Plesset equation (see, e.g., References and
36
individual cavity dynamics behavior, these turbulent fluctuations ), which is beyond the scope of the present investigation. The pres-
remain orders of magnitudes higher than other anticipated tran- ented CFD results do however offer a means to link cavity dynamics
sient mechanisms, namely vortex precession, which as reported predictions at the microscale to realistic, turbulent flow data at reactor
elsewhere is expected to be of the order of 60 Hz.24 The computed scale. This offers a route to optimize reactor designs with a view to
2D flow fields therefore form a suitable basis on which to perform creating the ideal conditions for maximum cavitation yield and cavity

F I G U R E 1 5 (a) Turbulent kinetic energy, scale effect. (b) Turbulence eddy frequency, scale effect [Color figure can be viewed at
wileyonlinelibrary.com]
14 of 18 SIMPSON AND RANADE

FIGURE 16 Predicted cavity trajectories, Device A [Color figure can be viewed at wileyonlinelibrary.com]

collapse intensity. It is instructive therefore to understand the typical p0 = p + ρk: ð13Þ


instantaneous pressure histories of the generated cavities. In order to
study individual cavity trajectories, a series of lagrangian calculations The frequency of oscillation can then be directly related to the turbu-
were performed on the solved continuous flow fields for each device. lence eddy frequency, ω, using an expression of the following form:
The instantaneous pressures experienced by the tracked cavities can
p0 = p + ρk sin ðωtÞ: ð14Þ
be related to the mean flow and turbulence quantities; the amplitude
of the pressure fluctuations, and the instantaneous pressure, p0 can be The trajectories were treated as point masses with density equal to
described by the following relationship: the water vapor density (0.55 kg/m3), assumed from a reference
SIMPSON AND RANADE 15 of 18

FIGURE 17 Predicted cavity trajectories, Devices B, C and D [Color figure can be viewed at wileyonlinelibrary.com]

pressure of 800 Pa and a temperature of 295 K. Drag on the particles is also necessary. Full details of these relationships can be found in
is imposed through a spherical drag law. Owing to the presence of a Reference 33. Larger cavities will tend to travel toward the axis of the
strong radial pressure gradient, and the large density differences device, where they may accumulate and coalesce in the central core.
between the discrete phase and continuous phase in this case, the Smaller generated cavities on the other hand will tend to follow the
inclusion of additional virtual mass force and pressure gradient terms path of the liquid phase, and will thus orbit the vortex core. At the
16 of 18 SIMPSON AND RANADE

point of area increase in the axial port, the high turbulence predic- cavitating bubbles, or groups of bubbles. The simulation results pres-
tions, coupled with relatively strong reverse flow, will lead to unsteady ented here highlight a highly turbulent region where the fixed vortex
break-up of the vortex core, producing highly complex interactions core terminates in the axial port, and detailed flow visualization exper-
between the gas phase and local turbulent eddies. In the present work iments are planned to build our understanding of the precise nature
therefore some simplifying assumptions are required in order to set the of the two-phase flow in this region. These experiments are being car-
starting point from which to initiate single cavity calculations. In this ried out and will be published separately. The developed 2D approach
instance, using the solved eulerian multiphase results, the cavities were and models will also provide a sound and useful basis for extending
tracked from a starting point on surface of constant volume fraction the models to full 3D, time-resolved simulations of cavitating flows in
equal to 1, considered to be the center of the gas core. To investigate vortex devices.
cavity size effects, Initial calculations were performed for bubble sizes of
10, 100 and 200 μm, tracked up to a maximum residence time of 0.01 s.
4 | C O N CL U S I O N S
Predictions showed cavity size to have a negligible influence on the com-
puted trajectories over this range, and as such the median size of
Multiphase computational fluid dynamics models of vortex cavitation
100 μm was applied for subsequent calculations. Limited investigations
devices were developed which successfully reproduce experimental
of larger sizes showed that for bubbles of 500 μm and above, the cavities
trends in flow rate versus pressure drop. The resulting models were
become trapped within the reverse flow central core, resulting in incom-
subsequently used to investigate the cavitation inception conditions,
plete trajectory calculations.
and thereafter the development of the two-phase flow and turbulence
The resulting instantaneous pressure histories are presented in
fields at successively increasing flow rates. This particular vortex
Figure 16 for Device A at different throat velocities. The graphs in
device design exhibits two prominent flow features; first, the strong
each case show a selected sample of 10 trajectories in total, with the
swirling flow sets up a strong radial pressure gradient, creating a cavi-
initial starting positions of each trajectory evenly spaced in the axial
tating vortex core in the axial port. Second, this strong vortex is
direction along the surface of the gas core, and each terminating after
accompanied by a region of reversed flow along the axis of the device,
a total residence time of 0.01 s. The trajectory time–pressure histories
which persists through the full length of the axial port. The vortex
show two oscillation mechanisms; a large amplitude fluctuation which
device thus differs from linear flow devices in two fundamental
exceeds the difference in outlet pressure (p2) and vapor pressure (pv),
and superimposed on these large scale fluctuations a higher fre- respects; unlike orifice or venturi devices, cavitation is initiated in the

quency, lower amplitude oscillation due to the local turbulence field. fluid bulk as opposed to the solid surface of the restriction. As such,

Due to the high radial pressure gradients around the vortex core (illus- through the use of swirling flows it is possible to eliminate the damag-

trated in the contour plots in Figure 16), as the cavities swirl around ing effects of clogging and erosion. Second, the strong radial pressure

the core they are continuously moving radially from the low-pressure gradient in the cavitating region leads to individual cavities experienc-

core at pv into high pressure regions (of the order of 20× pv) as they ing pressure rises of over 2× the device back pressure, whereas in lin-

are advected through the axial port. At higher throat velocities, the ear flow devices the maximum pressure recovery experienced by

plots highlight a shift in frequency of the fluctuations in instantaneous cavities is limited to the device pressure recovery. Discrete cavity tra-
pressure; both the large amplitude fluctuations, and also the lower jectory calculations highlight the large amplitudes in pressure recovery
amplitude, turbulence fluctuations show a clear increase with increas- experienced by individual cavities in the micron size range. The results
ing throat velocities. also highlight the key differences in hydrodynamic behavior with
The instantaneous pressure fluctuations are presented at different device scale; principally that the turbulence fluctuation frequencies
device scales in Figure 17, each at a common throat velocity, reduce as the device size is increased. This may have an effect on the
ut = 3.0 m/s. The key trend shown in the predicted results is that the resulting cavity dynamics, and therefore the collapse intensities
frequency of both large scale and small-scale turbulence oscillations achieved. This warrants further investigation, which will form the
show an obvious decrease as the device scale increases, reflecting the basis of future work in this area. There is, therefore, promising poten-
reduction in turbulence eddy frequency with increasing device scale. tial to harness the fundamental hydraulic features of vortex devices
Also indicated by the time–pressure plots is an apparent reduction in revealed in this work to produce optimized device designs which can
amplitude of the instantaneous pressures with increasing scale (from realize the full potential of HC for use in full scale industrial processes.
approx. 20× pv for Device B, down to 12× pv for Device D). The key findings of the present study are summarized as follows:
The results presented here provide useful insights into key flow
characteristics of cavitating flow in vortex devices, and highlight the • The strong swirling flow is shown to generate a cavitating vortex
unique steep radial pressure gradients generated by such high swirl core, which is restricted to the fluid bulk. This uniquely differs from
flow devices. The results will also be useful to inform the design of cavitation in linear flow devices commonly adopted in the available
systematic experiments to generate empirical data on the velocity and literature, which tends to initiate and evolve on the surface of the
turbulence fields to act as a basis for further rigorous validation of flow restriction.
computational models. Future work will focus in particular on the • The radial pressure gradient in the cavitating region is shown to
complex interactions between the central gas core and individual dominate over the axial pressure gradient. Smaller cavities (of the
SIMPSON AND RANADE 17 of 18

order of 10 μm) therefore experience pressure recovery rates in 13. Suryawanshi PG, Bhandari VM, Sorokhaibam G, Ruparelia JP,
excess of 2× the overall device pressure recovery. Ranade VV. Solvent degradation studies using hydrodynamic cavita-
tion. Environ Prog Sustain Energy. 2018;37(1):295-304. https://doi.
• Cavity trajectories are predicted to flow around the vortex core as
org/10.1002/ep.12674.
they are advected through the axial port, continuously moving 14. Gaikwad V, Ranade VV. Disinfection of water using vortex diode as
from low pressure (pv) to high pressure (up to 20× pv) due to the hydrodynamic cavitation reactor. Asian J Chem. 2016;28(8):1867-
large local pressure gradients. These large amplitude fluctuations 1870.
15. Suryawanshi NB, Bhandari VM, Sorokhaibam LG, Ranade VVA. Non-
are combined with lower amplitude fluctuations due to local turbu-
catalytic deep desulphurization process using hydrodynamic cavitation.
lence quantities. Sci Rep. 2016;6(September):1-8. https://doi.org/10.1038/srep33021.
• Turbulence kinetic energy in the cavitating region is predicted to 16. Maddikeri GL, Gogate PR, Pandit AB. Intensified synthesis of biodie-
be a linear function of velocity at all device scales (scaling factors sel using hydrodynamic cavitation reactors based on the inter-
esterification of waste cooking oil. Fuel. 2014;137:285-292. https://
up to 8). Turbulent eddy frequencies were found to be inversely
doi.org/10.1016/j.fuel.2014.08.013.
proportional to device scale. 17. Kelkar MA, Gogate PR, Pandit AB. Intensification of esterification of
acids for synthesis of biodiesel using acoustic and hydrodynamic cavi-
tation. Ultrason Sonochem. 2008;15(3):188-194. https://doi.org/10.
1016/j.ultsonch.2007.04.003.
ORCID
18. Terán Hilares R, Ramos L, da Silva SS, Dragone G, Mussatto SI, dos SJC.
Hydrodynamic cavitation as a strategy to enhance the efficiency of lig-
Alister Simpson https://orcid.org/0000-0003-0855-4772
nocellulosic biomass pretreatment. Crit Rev Biotechnol. 2017;38(4):483-
Vivek V. Ranade https://orcid.org/0000-0003-0558-6971 493. https://doi.org/10.1080/07388551.2017.1369932.
19. Nakashima K, Ebi Y, Shibasaki-Kitakawa N, Soyama H, Yonemoto T.
Hydrodynamic cavitation reactor for efficient pretreatment of ligno-
RE FE R ENC E S cellulosic biomass. Ind Eng Chem Res. 2016;55(7):1866-1871. https://
doi.org/10.1021/acs.iecr.5b04375.
1. Hreiz R, Gentric C, Midoux N, Lainé R, Fünfschilling D. Hydrodynam- 20. Albanese L, Ciriminna R, Meneguzzo F, Pagliaro M. Beer-brewing
ics and velocity measurements in gas-liquid swirling flows in cylindri- powered by controlled hydrodynamic cavitation: theory and real-
cal cyclones. Chem Eng Res Des. 2014;92(11):2231-2246. https://doi. scale experiments. J Clean Prod. 2017;142:1457-1470. https://doi.
org/10.1016/j.cherd.2014.02.029. org/10.1016/j.jclepro.2016.11.162.
2. Hreiz R, Gentric C, Midoux N. Numerical investigation of swirling flow 21. Carpenter J, George S, Saharan VK. Low pressure hydrodynamic cavi-
in cylindrical cyclones. Chem Eng Res Des. 2011;89(12):2521-2539. tating device for producing highly stable oil in water emulsion: effect
https://doi.org/10.1016/j.cherd.2011.05.001. of geometry and cavitation number. Chem Eng Process Process Intensif.
3. De Wilde J, Broqueville DA. Rotating fluidized beds in a static geome- 2017;116(2016):97-104. https://doi.org/10.1016/j.cep.2017.02.013.
try: experimental proof of concept. AIChE J. 2007;53(4):793-810. 22. Simpson A, Ranade VV. Modelling of hydrodynamic cavitation with
https://doi.org/10.1002/aic. orifice: influence of different orifice designs. Chem Eng Res Des. 2018;
4. Volchkov EP, Dvornikov V, Lukashov VV, Abdrakhmanov RK. Investi- 136:698-711. https://doi.org/10.1016/j.cherd.2018.06.014.
gation of the flow in the cortex chamber with centrifugal fluidizing 23. Simpson A, Ranade VV. Modeling hydrodynamic cavitation in Venturi:
bed with and without combustion. Thermophys Aeromech. 2013;20(6): influence of Venturi configuration on inception and extent of cavitation.
663-668. AIChE J. 2019;65(1):421-433. https://doi.org/10.1002/aic.16411.
5. Niyogi K, Torregrosa MM, Marin GB, Heynderickx GJ, Shtern VN. On 24. Pandare A, Ranade VV. Flow in vortex diodes. Chem Eng Res Des.
the mechanisms of secondary flows in a gas vortex unit. AIChE J. 2015;102:274-285.
2018;64(5):1859-1873. https://doi.org/10.1002/aic.16087. 25. Yin J, Li J, Ma Y, Li H, Liu W, Wang D. Study on the air core formation
6. Moholkar VS, Senthil Kumar P, Pandit AB. Hydrodynamic cavitation of a gas–liquid separator. J Fluids Eng. 2015;137(9):091301. https://
for sonochemical effects. Ultrason Sonochem. 1999;6(1–2):53-65. doi.org/10.1115/1.4030198.
https://doi.org/10.1016/S1350-4177(98)00030-3. 26. Arndt REA, Arakeri VH, Higuchi H. Some observations of tip-vortex
7. Suslick KS, Mdleleni MM, Ries JT. Chemistry induced by hydrody- cavitation. J Fluid Mech. 1991;229:269-289. https://doi.org/10.1017/
namic cavitation. J Am Chem Soc. 1997;119(39):9303-9304. https:// S0022112091003026.
doi.org/10.1021/ja972171i. 27. Arndt R, Pennings P, Bosschers J, van Terwisga T. The singing vortex.
8. Rajoriya S, Carpenter J, Saharan VK, Pandit AB. Hydrodynamic cavita- Interface Focus. 2015;5(5):1-11. https://doi.org/10.1098/rsfs.2015.
tion: an advanced oxidation process for the degradation of bio- 0025.
refractory pollutants. Rev Chem Eng. 2016;32(4):379-411. https://doi. 28. Ma J, Hsiao C-T, Chahine GL. A physics based multiscale modeling of
org/10.1515/revce-2015-0075. cavitating flows. Comput Fluids. 2017;145:68-84. https://doi.org/10.
9. Dular M, Griessler-Bulc T, Gutierrez-Aguirre I, et al. Use of hydrody- 1016/j.compfluid.2016.12.010.
namic cavitation in (waste)water treatment. Ultrason Sonochem. 2016; 29. Chen L, Zhang L, Peng X, Shao X. Influence of water quality on the tip
29:577-588. https://doi.org/10.1016/j.ultsonch.2015.10.010. vortex cavitation inception. Phys Fluids. 2019;31(2):023303. https://
10. Ranade VV, Bhandari VM. Industrial Wastewater Treatment, Recycling, doi.org/10.1063/1.5053930.
and Reuse-Past, Present and Future. Oxford, UK: Butterworth- 30. Wilcox DC. Turbulence Modeling for CFD. La Canada, CA: DCW Indus-
Heinemann Elsevier; 2014. tries; 1998.
11. Kalumuck KM, Chahine GL. The use of cavitating jets to oxidize organic 31. Menter FR. Two-equation eddy-viscosity turbulence models for engi-
compounds in water. J Fluids Eng. 2000;122:465-470. https://doi.org/ neering applications. AIAA J. 1994;32(8):1598-1605.
10.1115/1.1286993. 32. Menter FR, Kuntz M, Langtry R. Ten years of industrial experience
12. Capocelli M, Prisciandaro M, Lancia A, Musmarra D. Hydrodynamic cavi- with the SST turbulence model.Pdf. Turbul Heat Mass Transf. 2003;4(1):
tation of p-nitrophenol: a theoretical and experimental insight. Chem Eng 625-632.
J. 2014;254:1-8. https://doi.org/10.1016/j.cej.2014.05.102. 33. Ansys. FLUENT 17.0 Theory Guide. Ansys Inc.; 2016.
18 of 18 SIMPSON AND RANADE

34. Kato M, Launder B. The modeling of turbulent flow around stationary 51. Chen Y, Lu CJ, Wu L. Modelling and computation of unsteady turbu-
and vibrating square cylinders. Proceedings of the 9th Symposium on lent cavitation flows. J Hydrodyn. 2006;18(5):559-566. https://doi.
Turbulent Shear Flows, Kyoto, August; Springer Berlin-Heideberg. org/10.1016/S1001-6058(06)60135-2.
1993:10.4.1-10.4.6. https://doi.org/10.1007/s13398-014-0173-7.2. 52. Li D-Q, Grekula M, Lindell P. A modified SST k-ω turbulence model to
35. Brennan CE. Cavitation and Bubble Dynamics. New York: Oxford Uni- predict the steady and unsteady sheet cavitation on 2D and 3D
versity Press; 1995. hydrofoils. Int Symp Cavitation. 2009;107:1-13.
36. Alehossein H, Qin Z. Numerical analysis of Rayleigh-Plesset equation 53. Goncalvs E. Numerical study of unsteady turbulent cavitating flows.
for cavitating water jets. Int J Numer Meth Eng. 2007;72:780-807. Eur J Mech B/Fluids. 2011;30(1):26-40. https://doi.org/10.1016/j.
37. Sauer J, Schnerr GH. Unsteady cavitating flow: a new cavitation model euromechflu.2010.08.002.
based on a modified front capturing method and bubble dynamics. 54. Decaix J, Goncalves E. Time-dependent simulation of cavitating flow
FEDSM'00, ASME Fluids Engineering Summer Conference; American Soci- with k-l turbulence models. Int J Numer Methods Fluids. 2011;68:
ety of Mechanical Engineers, New York, NY. 2000:11-15. 1053-1072. https://doi.org/10.1002/fld.2601.
38. Zwart PJ, Gerber AG, Belamri T. A two-phase flow model for 55. Ji B, Luo X, Arndt REA, Wu Y. Numerical simulation of three dimen-
predicting cavitation dynamics. Fifth International Conference on sional cavitation shedding dynamics with special emphasis on
Multiphase Flow, Yokohama, Tsukuba, Japan; 2004. cavitation-vortex interaction. Ocean Eng. 2014;87:64-77. https://doi.
39. Singhal AK, Athavale MM, Li H, Jiang Y. Mathematical basis and vali- org/10.1016/j.oceaneng.2014.05.005.
dation of the full cavitation model. J Fluids Eng. 2002;124(3):617. 56. Long X, Cheng H, Ji B, Arndt REA, Peng X. Large eddy simulation and
https://doi.org/10.1115/1.1486223. Euler–Lagrangian coupling investigation of the transient cavitating
40. Battistoni M, Duke DJ, Swantek AB, Tilocco FZ, Powell CF, Som S. turbulent flow around a twisted hydrofoil. Int J Multiph Flow. 2018;
Effects of noncondensable gas on cavitating nozzles. At Sprays. 2015; 100:41-56. https://doi.org/10.1016/j.ijmultiphaseflow.2017.12.002.
25(6):453-483. https://doi.org/10.1615/atomizspr.2015011076. 57. Zaman E, Vakil A, Martinez M, Olson J. An integral criterion for turbu-
41. Freudigmann H-A, Dörr A, Iben U, Pelz PF. Modeling of cavitation- lence suppression in swirling flows. Can J Chem Eng. 2018;96:2025-
induced air release phenomena in micro-orifice flows. J Fluids Eng. 2034. https://doi.org/10.1002/cjce.23145.
2017;139(11):111301. https://doi.org/10.1115/1.4037048. 58. Bardina JE, Huang PG, Coakley TJ. Turbulence modelling validation.
42. Arndt REA, Keller AP. Water quality effects on cavitation inception in AIAA-1997-2121. 28th Fluid Dynamics Conference, Snowmass Village,
a trailing vortex. J Fluids Eng. 1992;114(3):430-438. https://doi.org/ CO; American Institute of Aeronautics and Astronics. 1997.
10.1115/1.2910049. 59. Derksen JJ, Van Den Akker HEA. Simulation of vortex core preces-
43. Jiao L, Zhang PP, Chen CN, Yin JL, Wang LQ. Experimental study on sion in a reverse-flow cyclone. AIChE J. 2000;46(7):1317-1331.
the cavitation of vortex diode based on CFD. IOP Conf Ser Earth Envi- 60. Dinesh KKJR, Kirkpatrick MP. Computers & fluids study of jet preces-
ron Sci. 2012;15(pt 6):062058. https://doi.org/10.1088/1755-1315/ sion, recirculation and vortex breakdown in turbulent swirling jets
15/6/062058. using LES. Comput Fluids. 2009;38(6):1232-1242. https://doi.org/10.
44. Reboud J-L, Stutz B, Coutier O. Two-phase flow structure of cavita- 1016/j.compfluid.2008.11.015.
tion: experiment and modelling of unsteady effects. 3rd International 61. Brinkhorst S, von Lavante E, Wendt G. Experimental and numerical
Symposium on Cavitation CAV1998, Université Joseph Fourier, Grenoble, investigation of the cavitation-induced choked flow in a herschel
France; Université Joseph Fourier; 1998:203-208. venturi-tube. Flow Meas Instrum. 2017;54(November 2016):56-67.
45. Coutier-Delgosha O, Fortes-Patella R, Reboud JL. Evaluation of the tur- https://doi.org/10.1016/j.flowmeasinst.2016.12.006.
bulence model influence on the numerical simulations of unsteady cavita- 62. Ashrafizadeh SM, Ghassemi H. Experimental and numerical investiga-
tion. J Fluids Eng. 2003;125(1):38. https://doi.org/10.1115/1.1524584. tion on the performance of small-sized cavitating venturis. Flow Meas
46. Charriere B, Decaix J, Goncalves E. A comparative study of cavitation Instrum. 2015;42:6-15. https://doi.org/10.1016/j.flowmeasinst.2014.
models in a Venturi flow. Eur J Mech B/Fluids. 2015;49(PA):287-297. 12.007.
https://doi.org/10.1016/j.euromechflu.2014.10.003. 63. Soyama H, Hoshino J. Enhancing the aggressive intensity of hydrody-
47. Charrière B, Goncalves E. Numerical investigation of periodic cavita- namic cavitation through a Venturi tube by increasing the pressure in
tion shedding in a Venturi. Int J Heat Fluid Flow. 2017;64:41-54. the region where the bubbles collapse. AIP Adv. 2016;6(4):1–12.
https://doi.org/10.1016/j.ijheatfluidflow.2017.01.011. https://doi.org/10.1063/1.4947572.
48. Ji B, Luo X, Wu Y, Peng X, Duan Y. Numerical analysis of unsteady 64. Capocelli M, Prisciandaro M, Lancia A, Musmarra D. Cavitational reac-
cavitating turbulent flow and shedding horse-shoe vortex structure tor for advanced treatment of contaminated water: the effect of
around a twisted hydrofoil. Int J Multiph Flow. 2013;51:33-43. recovery pressure. Desalin Water Treat. 2015;55(12):3172-3177.
https://doi.org/10.1016/j.ijmultiphaseflow.2012.11.008. https://doi.org/10.1080/19443994.2014.947779.
49. Zhang DS, Wang HY, Shi WD, Zhang GJ, Van Esch BPM. Numerical
analysis of the unsteady behavior of cloud cavitation around a hydro-
foil based on an improved filter-based model. J Hydrodyn. 2015;27(5):
795-808. https://doi.org/10.1016/S1001-6058(15)60541-8. How to cite this article: Simpson A, Ranade VV. Flow
50. Long X, Cheng H, Ji B, Arndt REA. Numerical investigation of attached
characteristics of vortex based cavitation devices. AIChE J.
cavitation shedding dynamics around the Clark-Y hydrofoil with the
FBDCM and an integral method. Ocean Eng. 2017;137(December 2019;e16675. https://doi.org/10.1002/aic.16675
2016):247-261. https://doi.org/10.1016/j.oceaneng.2017.03.054.

Potrebbero piacerti anche