You are on page 1of 17

Ultrasonics Sonochemistry 106 (2024) 106875

Contents lists available at ScienceDirect

Ultrasonics Sonochemistry
journal homepage: www.elsevier.com/locate/ultson

Cavitation suppression and transformation of turbulence structure in the


cross flow around a circular cylinder: Surface morphology and
wettability effects
Mikhail Yu. Nichik a, Boris B. Ilyushin a, Ebrahim Kadivar b, *, Ould el Moctar b, Konstantin
S. Pervunin a, *
a
Kutateladze Institute of Thermophysics, Siberian Branch of the Russian Academy of Sciences (IT SB RAS), 630090 Novosibirsk, Russia
b
Institute of Ship Technology, Ocean Engineering and Transport Systems, University of Duisburg-Essen, 47057 Duisburg, Germany

A R T I C L E I N F O A B S T R A C T

Keywords: Passive methods of flow and cavitation control appear to offer some of the best prospects in the field of hydraulic
Cavitation control engineering and marine applications. In this article, we aimed at an experimental examination of the effect of
Wall roughness wall roughness/wettability on the occurrence of cavitation and turbulence structure in the cross flow around and
Wettability
in the wake of a circular cylinder in two characteristic regimes. For this, we used three test bodies with different
Visualization
surface morphologies: smooth (reference), micro-scale irregularities (rough) and regular large-scale (of the order
PIV
Statistical filtration of a millimeter) texture (finned). Using high-speed imaging to observe vapor cavities, we revealed that cavitation
Vapor fraction is noticeably suppressed by both types of roughness. Applying the method of vapor phase detection (Pervunin
Higher-order moments et al., 2021), this finding was then quantitatively confirmed through an in-depth analysis of an ensemble of
Bimodal distributions instantaneous velocity fields measured by PIV, indicating that modification of wall morphology is an effective
method of cavitation control. The procedure of statistical vector filtration (Heinz et al., 2004) allowed us to
remove outliers from the velocity fields and, thus, calculate various turbulence characteristics, including higher-
order moments (i.e., the coefficients of skewness and excess). Wall irregularities were found to significantly affect
the turbulence structure of the wake flow, but the higher-order moments downstream of the modified-surface
cylinders turned out to be unexpectedly insensitive to a change in the flow regime, as opposed to the smooth
one. Regardless of the type of surface morphology, the influence of roughness on the mechanism of formation of
large-scale vortices and their characteristics was weakened. However, it caused overall disorganization of liquid
motion in the cylinder wake, thus making local flow conditions highly unsteady. In addition, this process became
more chaotic with an increase in the scale of irregularities.

1. Introduction passage of an acoustic wave during the half-periods of its expansion


(ultrasound cavitation), etc. Various factors like turbulent flow charac­
Hydrodynamic cavitation is often a harmful phenomenon in hy­ teristics [65,69,70], surface roughness [3], temperature and viscosity of
draulic engineering systems and marine applications that usually occurs the carrier liquid [30], nuclei density and radius [68] and the size of
on propeller blades, ship rudders and moving parts of hydraulic systems. vapor nanobubbles [23,24] affect cavitation inception and develop­
It causes such undesirable effects as material erosion, high vibrational ment. Moreover, these parameters may also influence each other in
loads, increased noise and overall reduction of hydrodynamic perfor­ complex ways. At present, the mechanism of interaction of wall irreg­
mance (e.g., [54,19,27,51,36,37,56]. It is well known [21] that cavita­ ularities with turbulence and, especially, large-scale vortex structures is
tion is initiated on immersible bodies in the flow regions where pressure not yet fully understood, for both streamlined and bluff bodies.
locally falls below the one of saturated vapor of the surrounding liquid at The literature is currently abundant in studies on a round cylinder in
a constant temperature, which can, for example, occur because of an crossflow. Various aspects of this problem have been examined so far,
increase in the flow velocity, a sudden change in the duct geometry, the such as pressure distribution, vortex formation and overall flow pattern,

* Corresponding authors.
E-mail addresses: ebrahim.kadivar@uni-due.de (E. Kadivar), konstantin.pervunin@gmail.com (K.S. Pervunin).

https://doi.org/10.1016/j.ultsonch.2024.106875
Received 15 February 2024; Received in revised form 31 March 2024; Accepted 10 April 2024
Available online 11 April 2024
1350-4177/© 2024 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC license (http://creativecommons.org/licenses/by-
nc/4.0/).
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

forces and torque acting on the test object, etc. For engineers, the main boundary layer separation under an adverse pressure gradient, also
practical interest was the average drag and lift coefficients under reducing the drag.
different flow conditions. In the book by Zdravkovich [75], a functional Hence, liquid flow over a solid wall strongly depends both on its
dependence of the drag on the Reynolds number was established and, on material wettability (basically, hydrophility/hydrophobicity), which
its basis, a classification of five characteristic regimes of the transverse affects the development of the boundary layer, as indicated above, and
flow around a cylinder was proposed. Achenbach [1] and Güven et al. therefore the structure of its turbulent wake, and on the occurrence of
[26] provided experimental data on the drag coefficient for similar test cavitation. The wetting properties are, in turn, determined by the
objects with rough surfaces. For symmetric bodies like a sphere or cyl­ following two factors: roughness and the level of surface energy
inder with identical conditions above and below relative to the flow [35,60,42]. Hydrophobic materials have unique features, including self-
direction, the effect of roughness on the lift force is obviously negligible. cleaning, protection against corrosion, icing and fogging, etc. [72],
However, shifting the boundary layer separation point to the rear of an which are widely found in nature, including plant leaves [6], wings
object leads to normalization of pressure distribution over its surface [50,17] and legs [22] of insects and animals [5]. To date, artificial hy­
and, thus, to a decrease in the drag. As demonstrated by Tian et al. [64] drophobic coatings have been also developed and used in practice,
and Zhang et al. [76], the mechanism of drag reduction in cavitating which are based on low-energy surfaces and nano- or micro-scale
flows is associated with the combined influence of both modification of roughness (e.g., [18,73,59]). At present, the mechanisms of interac­
their turbulence structure and phase slip on the cavity interface. Large- tion of hydrophobic materials with liquid flows are actively studied (e.
scale vortices are forced to move out of the boundary layer, as if they g., see the reviews [47] and [62]). In fluid dynamics, one of the most
were ‘raised’ above the surface [64,76], with the flow velocity profile important advantages of using hydrophobic coatings is also a decrease in
becoming flatter. According to Malkin et al. [46], hydrophobicity also drag of underwater bodies [25,32].
enhances slip on a solid surface. Thus, an attached vapor cavity and a Brandner et al. [10] discussed in detail the mechanism of cavitation
hydrophobic wall both facilitate flattening of the flow velocity profile, inception on a metallic sphere in the form of individual vapor streaks of
which implies that the nature of the effects of surface hydrophobicity approximately the same size distributed almost uniformly around the
and roughness on vapor formation and drag reduction are apparently entire circumference. The streaks present continuously at certain loca­
similar and associated with liquid slip on the wall. It is worth noting that tions grow and merge to establish an attached cavity with a cellular-
the number of articles dedicated to the effect of surface wettability on structured leading edge. Before then, Leger and Ceccio [43] also
cavitating flow is scarce. observed a similar cavitation pattern on a hydrophilic sphere, which
So, irregularities on the surface of a body immersed in a fluid flow was, however, absent on a hydrophobic one. Note that this cellular
can either reduce or increase its drag. Despite extensive advances in the structure of the incipient cavity resembles the process of formation of
fundamental understanding of turbulence generation over rough walls, Λ-structures [9] and Theodorsen horseshoe-shaped vortices [2] at the
this issue is still far from being fully resolved (e.g., see the review by initial stage of evolution of turbulent boundary layer. At a near-critical
Kadivar et al. [38]). The reasons lie in the chaotic and random nature of cavitation number, these vortices create low-pressure areas in a liquid
turbulent flows, the lack of systematic studies of their structure and the flow, where the phase transition must occur in the first place, presum­
variety of types of roughness that can potentially significantly influence ably producing the above-mentioned cellular structure. Since the
their dynamics. For example, Jouybari et al. [34] proposed a polynomial horseshoe vortices arise from the flow slowdown by the wall, a decrease
model of surface topography with as many as 30 parameters (degrees of in the deceleration rate or, in other words, an increase in the slip length
freedom) which efficiency was verified on 45 kinds of rough plates. leads to a delay in their formation or their complete elimination in the
Furthermore, the mechanisms of interaction of fluid flow with individ­ case of insufficient extension of the surface. Boiko et al. [9] indicated
ual irregularities can be also substantially different in liquids compared that the transition to turbulence occurs much faster on a smooth plate
to gases because of wall wetting and phase transitions like vaporization than a ribbed one. Thus, an increase in the slip length on a hydrophobic
and cavitation. In particular, roughness can either enhance or reduce surface or, which is equivalent, a rise in the flow velocity at the wall
fluid slip [46], locally increasing or decreasing the slope angle of the should result in retardation of the development of the boundary layer,
mean velocity profile near the surface, which leads either to a delay in less pronounced deformation of the velocity profile and, as a conse­
flow separation or turbulization of the boundary layer and, on the quence, smaller number of the low-pressure areas.
contrary, its faster stall. This in turn affects the hydrodynamic quality of Various passive methods of cavitation control based on these prin­
the test body. For example, flow separation on a hydrofoil is accompa­ ciples were discussed in detail in the review by Zaresharif et al. [74],
nied by the formation of reverse fluid motion and, as a result, a sharp including those employing the concept of variation of the contact angle.
drop in the lift and a rise in the drag. Numerous papers have been published so far, where the effect of wall
The impact of small- and large-scale roughness on the near-wall flow irregularities on flow structure was investigated in detail for both a
is different. When liquid is in full contact with a rough surface (i.e., no cylinder (e.g., [71,63,55,8,77,45]) and other objects of different shapes
gas or vapor inclusions are present within its recesses), which is referred (e.g., [53,39,7,33,11,34]). In particular, Arndt and Ippen [4], Coutier-
to as the Wenzel state (e.g., see [47]), small-scale roughness prevents the Delgosha et al. [14], Li et al. [44], Danlos et al. [15] and Churkin
liquid from slipping thanks to additional disturbances in the velocity et al. [11] examined the roughness effect on cavitation, which is directly
field induced by irregularities [46]. The larger the average height/depth related to the surface wettability, confirming that the interaction of
of bumps is, the more pronounced this effect becomes. However, the large- and small-scale irregularities with vapor cavities occurs in
build-up of gas/vapor in the surface recesses enhances the slip effect different ways. van Rijsbergen [67], in turn, provided a detailed review
[46] due to the alternating contact of the liquid moving along the wall on the physical mechanisms of cavitation inception and how roughness
with roughness protrusions and the dispersed phase held in between, affects them. Leger and Ceccio [43] experimentally demonstrated that
thus making the surface hydrophobic or superhydrophobic if the ma­ the physical properties of the wall material significantly influence the
terial itself is hydrophobic, which is known as the Cassie–Baxter state shape of an attached cavity and the location of boundary layer separa­
[47]. In which case, the drag is obviously less (the slip is, on the con­ tion in both single-phase and cavitating flows around a hydrophilic brass
trary, higher) than for a smooth solid wall as a result of reduced friction and hydrophobic PTFE sphere, as well as a cylinder. However, some
of the liquid against the formed gas/vapor cushions. For large-scale papers reporting on modification of the flow structure by hydrophobic
roughness, if the height of irregularities is greater than the thickness and rough surfaces contain conflicting results [13,11,38]. Thus, further
of the viscous sublayer, a recirculating flow occurs in the troughs (e.g., research in this direction is of great importance.
[38]), which results in a decrease of the drag coefficient [12]. In some In this article, we experimentally investigate the effect of different
cases, streamwise vortices are induced [40], which effectively delay the types of surface roughness (regular large-scale and irregular small-scale)

2
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

on the development of cavitation and the structure of turbulence in the diameter D = 25 mm and length (span) of 80 mm. Each of them
wake of a round cylinder in two characteristic regimes through high- expanded over the entire width of the test channel and was tightly
speed visualization of vapor cavities and Particle Image Velocimetry attached to both sidewalls. Two of these bodies were made of a brass
(PIV) measurements of velocity fields. Using the procedure of statistical round rod, while the third one was manufactured from a copper tube
filtration of velocity vectors [28] and the method of vapor phase with a 2-mm thick wall tightly fitted on a 21-mm diameter support. Each
detection [52], we obtain spatial distributions of the first four statistical of the cylinders was subject to a specific surface treatment. One with a
moments of the flow velocity and the time-averaged vapor concentra­ relatively smooth wall was produced by CNC milling and used as a
tion. The layout of this paper is as follows. In the next section, we give a reference (Fig. 1a). The second was coated with 46 g (1150 shots) of
detailed description of the test bodies, flow conditions and measurement nichrome powder (Cr20Ni80, fraction 1–40 μm), thus creating a solid
techniques. Section 3 is the primary one, where main results of the study layer of 250–280 μm thickness, in the Lavrentyev Institute of Hydro­
are presented and discussed. In Section 4, we summarize our findings dynamics SB RAS by detonation spraying [66] to have irregular
and draw conclusions. morphology of micro-scale roughness (Fig. 1b). The last one was
machined on a lathe in the Bauman Moscow State Technical University
2. Experiment through deformation cutting (spindle speed = 1600 rpm, minimum wall
thickness = 1.5 mm, maximum fin height = 3 mm, fin spacing =
2.1. Test bodies and surface morphology 0.25–1.5 mm) [61,78] to produce a regular texture of equidistantly-
spaced millimeter-sized fins of a spiral thread (Fig. 1c).
The experiments were conducted in the cavitation tunnel in Institute Roughness characteristics of the test bodies were measured with a
of Thermophysics SB RAS which description along with the monitoring portable profilometer Mahr GmbH MarSurf XR 1 equipped with a drive
and controlling equipment is available in Kravtsova et al. [41]. The test unit MarSurf SD 26, a skidless probe system BFW 250 and a BFW probe
section of the experimental facility is a 1.3-m long rectangular channel, arm A 10–45-2/90, where ‘10’ is the height of the stylus tip in milli­
with the inner height H = 250 mm and width W = 80 mm (more details meters, ‘45’ stands for the probe element offset relative to the pivot
can be found in [58]). The test objects are three cylinders of effective point at the x-axis in millimeters, ‘2’ denotes the stylus tip radius in

Fig. 1. Visual appearance (top row), close-up view (middle row) and characteristic surface profiles (bottom row) of the (a) smooth, (b) rough and (c) fin­
ned cylinders.

3
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

microns, ‘90’ reflects the cone angle of the stylus tip. This instrument specifically to measure the contact angle between immiscible fluids,
operates according to standard ISO 3274:1996 to measure a wall profile where one of them appears as a blob. The contact angle between the
and calculate its roughness parameters. The device readings are recor­ sample and liquid was quantified with an error of 0.1◦ through the
ded and processed in the MahrWin EasyRoughness software. The tangent method built in the same software [48]. For the smooth and
measured surface profiles were smoothed with a Gaussian filter in rough cylinders, 2 μL of water at a temperature of 20 ◦ C was placed on
accordance with ISO 16610–21 (cutoff wavelength = 0.8 mm). The main the tested sample to create a sessile drop. On the finned surface, a
roughness characteristics according to standards ISO 4287:1997 droplet of 10 μL volume put on the top of the thread was unstable and,
(roughness parameters), ISO 4288:1996 (measuring conditions) and ISO with a slight vibration, immediately leaked into the troughs. In the latter
11562:1998 (roughness/waviness filter) are the arithmetical mean de­ case, its volume was then also reduced to 2 μL to observe its shape be­
viation (level) Ra , maximum peak-to-valley height Rz , maximum peak tween the ribs.
height Rp , maximum valley depth Rv , mean spacing of peaks or valleys In our experiment, the materials of the cylinder surfaces, brass,
RSm , profile skewness RSk , its kurtosis RKu and root-mean-square slope copper and nichrome, are rather hydrophilic. Another supplementary
Rdq (see Table 1). For the smooth cylinder, RSm equals exactly 100 μm as test on thin flat foils of the same metals/alloys using the above method
it corresponds to the cutting tool step. It can be seen in Fig. 1a that the applied to a 2 μL water drop showed that the contact angle for brass,
average level of roughness of the smooth surface is low and the span copper and nichrome makes respectively about 88◦ , 80◦ and 76◦ . This is
between its peaks and troughs is insignificant. The wall of the rough also confirmed by Fig. 2a, where the contact angle of the smooth brass
cylinder is chaotically micro-structured in terms of the location of ir­
regularities and their size (Fig. 1b). The surface texture of the finned
cylinder is featured by alternating skewed ribs of height h ≈ 1.1 mm and
0.36-mm length and flat grooves of 0.64-mm length in the axial (span­
wise) direction which period (step) equals p = 1 mm (Fig. 1c). Thus, the
scales of the two types of roughness differ by two orders of magnitude.
In the case where a rough surface is in contact with a flowing liquid,
it is evident that the most important effect of irregularities is a modifi­
cation of the mean velocity profile near the wall (see Introduction)
which ultimately affects the development of boundary layer, its struc­
ture, conditions for the occurrence of cavitation, etc. The Reynolds
number of roughness is a quantitative characteristic that determines the
degree of its interaction with the buffer layer of the flow. Irregularities
disturb this layer but do not practically transfer anything to the liquid
bulk, except for a change in the surface friction drag. However, we
should acknowledge that Jiménez [33], for example, indicated a stron­
ger influence of roughness on the external flow. As mentioned above,
roughness of a wall affects its hydrophobic properties which are man­
ifested by the presence, volume and shape of a gas/vapor layer between
the liquid and solid surface, as well as between individual irregularities.
This, in turn, leads to an alleviation in the effect of the rough wall both
on the buffer layer and on the flow as a whole. If the cavitation number is
close to a critical value at which cavitation is initiated, roughness ele­
ments can also act as cavitation nuclei, thereby facilitating the genera­
tion of vapor and leading to a hydrophobicity-like effect, due to the
formation of local areas of reduced pressure even in the case of a hy­
drophilic surface [4].

2.2. Surface wettability

In order to examine the roughness effect on the wall wettability, we


took supplementary measurements of the static contact angle for each of
the three test bodies with different surface morphology. For this, we
sequentially fulfilled the following procedures. Immediately before the
test, all the cylinders were cleaned and degreased with ethanol. After
installing a test sample on a linear translation stage, the wall level, angle
of inclination and curvature of the former were determined manually
and saved in the DSA3 software. In the test, a drop of water was placed Fig. 2. Photographs of a water drop resting on the (a) smooth and (b) rough
on the sample, which volume was selected individually for each surface cylinders. The blue horizontal or close-to-horizontal line illustrates the surface
level. The image parts lying under this line are reflections of the upper part and,
in the range from 1 to 10 μL depending on its morphology. Shadowgraph
therefore, do not convey any useful information. The other blue curves depict
pictures of the drop were taken perpendicular to the cylinder axis using a
reconstructed imaginary arcs of the drop boundary needed to draw the tangent
multifunction system Kruss DSA-100E (range of contact angles = (green solid line) at its base.
0–180◦ , maximum sample dimensions = 300 × 150 mm) designed

Table 1
Main parameters characterizing the properties of roughness for the smooth and rough surfaces measured over a traversing length of 5.6 mm.
parameter Ra , μm Rz , μm Rp , μm Rv , μm RSm , μm RSk RKu Rdq

smooth 1.621 11.136 4.139 7 100 –0.787 4.132 0.282


rough 6.122 33.354 16.133 17.221 211.706 0.072 3.137 0.538

4
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

wall lies in the range of 82.2–84.0◦ . The difference in the contact angle 2.3. Flow conditions and measurement techniques
values for the brass cylinder and foil may be due to their different de­
grees of oxidation as the former was in contact with water for a long In this study, we consider two regimes of the cross flow around the
time. As for the micro-scale irregular roughness, the contact angle of this cylinders of different surface morphology as given in Table 2. In order to
surface is much higher and equals 128.6–130.0◦ (Fig. 2b). Thus, it ex­ change the flow conditions, we varied the Reynolds and cavitation
hibits stable and pronounced hydrophobic properties, though both the numbers together by adjusting the freestream bulk flow velocity U0 ,
brass substrate and nickel–chromium grains are individually hydro­ which are respectively defined as follows:
philic. Meanwhile, on the cylinder with the large-scale regular struc­
DU0 pin − pV
tures, a drop of water was initially located on the top of the thread, Re = , σc = ,
ν ρU02 /2
without penetrating into the grooves between each pair of the fins
(Fig. 3a). The wetting behavior of a liquid drop on a similar ribbed where pin is the static pressure quantified by a diaphragm strain-gauge
surface was described in detail by Manoharan and Bhattacharya [47] for pressure transducer at the test channel sidewall near its inlet with an
different damping conditions. However, this state of the system is error of ±1.5 kPa, pV = 4.24 kPa [16] is the saturation vapor pressure of
metastable: with weak shaking or increasing the drop volume, water the working liquid (distilled water at a temperature of 30 ◦ C in the
slides down to fill the free spaces between the roughness elements, present experiment), ρ and ν are the density and kinematic viscosity of
thereby distinctly revealing hydrophilic properties of the structured the working liquid.
surface (Fig. 3b). It is obvious that, if such a cylinder is completely U0 was measured far upstream of the location of the test body by PIV
immersed in a water flow, the liquid should fill the entire volume be­ in the central vertical longitudinal section over the entire height of the
tween the ribs. Hence, the presence of gas and/or water vapor inside test channel, which had a top-hat velocity profile, with the initial level of
these millimeter-scaled channels can just be caused by the phase tran­ turbulent fluctuations not exceeding 1.5 % of U0 at the test channel inlet.
sition (vaporization) in the resulting local zones of low pressure and the Table 2 also contains additional properties needed to fully describe the
release of dissolved air. flow conditions: pout is the static pressure at the outlet of the test channel
determined in the same way as pin , f is the frequency of generation of
cavitating Karman vortices in the wake of the cylinder that is evaluated
from high-speed visualization, and St = fD/U0 is the reduced frequency
(Strouhal number) of the Karman vortex street. The concentration of
dissolved oxygen in the working liquid was monitored by a Mettler
Toledo InPro 6850i sensor (measurement range from 6 ppb to satura­
tion, relative error 1 %) connected to a M400 multi-parameter trans­
mitter and maintained almost constant at about 7.4 mg/L.
High-speed imaging was performed both from the top and from the
side using a Photron FASTCAM SA5 camera (maximum resolution =
1024 × 1024 pixels, maximum framing rate = 7 kHz at full resolution)
equipped with a Nikon AF Micro Nikkor 60 mm f/2.8D lens. It was
mounted above or at the side of the test section, such that the distance
between the camera matrix and the nearest channel wall equaled 400
and 512 mm in case of the vertical and horizontal arrangements of the
visualization system. In the experiment, the image acquisition frequency
was set to 20 kHz at a frame size of 768 × 480 (top view) and 960 × 392
(side view) pixels. The optical magnification made up 0.12 and 0.18,
while the spatial resolution was 6 and 8.76 pixels/mm respectively for
the vertical and horizontal setups. The overall length of recording of
each experimental run was 1 s (20,000 frames in total). To ensure suf­
ficient image exposure, we used two high-power matrix LEDs 3F100
(maximum power = 100 W, color temperature = 6000 K, opening angle
= 140◦ ), each equipped with an individual divergent lens, with the
working level of direct current of 2.5 A and voltage of 30 V, to illuminate
cavitation structures in the cross flow around the cylinders. The angle of
illumination and position of each LED relative to the test body was
adjusted independently in both vertical and horizontal directions.
In order to measure the flow velocity, Particle Image Velocimetry
was employed. For this, a system comprised of a double-cavity pulsed
Nd:YAG Quantel EVG00200 laser (wavelength = 532 nm, pulse repeti­
tion rate = 15 Hz, pulse duration = 10 ns, maximum pulse energy = 200
mJ), a CCD-camera ImperX Bobcat IGV-B4820 (digit capacity = 16 bits,
matrix resolution = 4904 × 3280 pixels, maximum framing rate = 4 Hz
at full resolution) equipped with a Nikon AF Nikkor 50 mm f/1.4D lens
and a POLIS pulse/delay generator was used. Management of this
measurement system was implemented through a computer with the
ActualFlow software, as well as PIV data storage and processing. In the
experiment, the recording rate was equal to 4 Hz at a resolution of 3024
Fig. 3. Photographs of a water drop lying (a) on the top of the ribs and (b) × 1800 pixels. The thickness of a laser light sheet produced by a sheet-
inside the grooves of the finned cylinder. The blue horizontal line illustrates the forming optical assembly to illuminate seeding particles suspended in
upper edge of the cylinder surface passing through the tips of the protrusions. the flow was approximately 0.8 mm in the measurement section that
The other blue curves depict reconstructed imaginary arcs of the drop boundary coincided with the central vertical longitudinal plane of the test channel.
needed to draw the tangent (green solid line) at its base.

5
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

Table 2
Main integral parameters and dimensionless criteria of similarity characterizing the flow regimes examined.
Regime Surface pin , kPa pout , kPa U0 , m/s Re, ×105 σc f, Hz St

R-I smooth 106 ± 1.5 108.3 ± 1.5 9.2 ± 0.18 2.9 ± 0.06 2.4 ± 0.1 189 0.25 ± 0.01
rough – –
finned – –

R-II smooth 105.8 ± 1.5 104.8 ± 1.5 9.9 ± 0.2 3.1 ± 0.06 2.1 ± 0.08 217 0.27 ± 0.01
rough 225 0.28 ± 0.01
finned 200 0.26 ± 0.01

The distance between the camera and the laser sheet equaled 566 mm. by high-speed imaging show a similar trend of the vapor cavity evolu­
The physical dimensions of the measurement region were 180 × 107 tion depending on the presence and scale of irregularities – cavitation
mm. In this arrangement of the measurement system, the optical development occurs to be significantly hampered on the rough and
magnification and spatial resolution of the camera were 0.12 and 16.75 finned cylinders (Fig. 4). At the same time, this effect is more pro­
pixels/mm, respectively. The processing and validation procedures of nounced (i.e., cavitation is even further suppressed) in the case of micro-
PIV data are fully described in Ilyushin et al. [31]. A detailed analysis of scale irregular roughness compared to the millimeter-scaled regular
the measurement uncertainties is available in Nichik et al. [49]. microstructuring (cf. Fig. 4b and 4c). This finding is, however, in
contradiction with the results by Churkin et al. [11] for a NACA0015
3. Results and discussion hydrofoil, which were obtained in the same cavitation tunnel.
Churkin et al. [11] found that both a growth of Ra and Rz and a
As noted above, along with the morphology and chemical composi­ reduction in RSm of the hydrofoil surface, whose irregular roughness was
tion of a solid wall, cavitation is also influenced by the dynamic char­ created at the same installation and by the same method of detonation
acteristics of a liquid flow which include but not limited to the cavitation spraying, generally lead to an elongation of the attached cavity and a
number, statistical properties of turbulence and noise level [4]. That is faster transition to unsteady flow conditions, regardless of the angle of
why, when analyzing the effect of the surface roughness of a test body on attack. This implies that individual micro-scale irregularities on the
cavitation by comparing the results of independent experimental runs, it surface of a streamlined body, like a hydrofoil, must induce relatively
is important to secure the identity of all experimental conditions, except strong streamwise vortices in the form of streaks that promote cavitation
for parameters of irregularities. According to Table 2, this requirement is in their cores. It looks especially plausible at the leading edge of a hy­
fully met here, including the same experimental facility, shape and size drofoil where the thickness of the boundary layer is extremely small
of the test body, liquid properties, flow conditions, measurement and [41], so that these irregularities protrude into the bulk of the main flow
data processing techniques, etc. above the boundary layer, making this method of passive flow control
ineffective on streamlined objects [29].
As for a bluff body, such as a cylinder, the thickness of the boundary
3.1. Visual observation of cavitation layer over its surface grows very quickly downstream of the 90◦ point
due to the significant curvature of its shape. A rise in the level of
Earlier studies (e.g., [4,14,44,15,67]) revealed that roughness on roughness of the solid wall in this instance must result in intense
average results in cavitation mitigation. In this research, snapshots taken

Fig. 4. Top-view photographs of vapor structures on the surface and in the wake of the cylinders at the moment of maximum cavitation intensity in regime R-II
(Table 2) for the three types of wall morphology (Fig. 1): (a) smooth; (b) rough and (c) finned.

6
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

turbulization of the boundary layer, thus delaying its separation (i.e.,


〈v4 〉
shifting its separation point farther downstream, closer to the wake axis) Ev = − 3 − excess coefficient (fourth moment),
[57]. This in turn disturbs the wake flow, violating its steady pattern σ4v
and, thereby, making local conditions unfavorable for cavitation to
where ̃ v is the instantaneous flow velocity and v is its fluctuating
occur [20]. This hypothetical mechanism should be responsible for this
component.
huge difference in the effectiveness of the considered method of passive
As seen in Fig. 5, the fields of the excess coefficient of velocity
cavitation control between the cases of a bluff and streamlined body.
fluctuations in the transverse direction Evy practically coincide, irre­
spective of the sample size (cf. Fig. 5a and 5b), which leads us to the
3.2. Statistical analysis of turbulence structure
conclusion that the ensemble volume of 1000 realizations is sufficient to
calculate the first four moments of turbulent fluctuations, provided that
In order to calculate central moments of turbulent fluctuations of the
the procedure of statistical filtration [28] is employed. The representa­
flow field, the procedure of statistical filtration of velocity vectors [28]
tiveness of the sample volume to evaluate the time-averaged vapor
was applied to the entire PIV dataset separately for either regime from
content using the method of quantification of the probability of
Table 2. This validation method was previously demonstrated by
dispersed phase occurrence [52] was verified in a similar fashion (see
Ilyushin et al. [31] to be an effective tool to remove vector outliers from
Fig. 6). Besides, to ensure the correctness of computation of the vapor
the ensemble of velocity fields, specifically in cavitating flow, thereby
fraction, we also checked the absence of correlation between successive
making calculation of higher-order statistical moments of turbulent
realizations in the obtained dataset of instantaneous flow velocity. As
fluctuations possible, which are known to be highly sensitive even to
visible in Fig. 7, each of the distributions obeys an exponential law,
single incorrect measurements. The representativeness of the statistical
which implies all events in the ensembles are statistically independent
ensembles of the flow velocity was checked comparing spatial distri­
[52].
butions of the higher-order moments of turbulent fluctuations calculated
As it comes from Figs. 8 and 9, the spatial distributions of the time-
over N = 1000 and 5000 realizations (i.e., single measurements of
averaged flow velocity are completely different downstream of the
instantaneous velocity fields) as shown in Fig. 5 by the example of the
smooth and rough cylinders, even for the streamwise mean velocity Vx
excess coefficient. The first four moments of velocity fluctuations are
in the main part of the turbulent wake (cf. Fig. 8a.1 and 8b.1 for R-I,
given by the following formulae resulting from the Reynolds decom­
Fig. 8a.2 and 8b.2 for R-II). This indicates a dramatic influence of the
position (̃
v = V + v):
small-scale irregular roughness on the flow structure. Meanwhile,
V = 〈̃
v〉 − time-averaged(mean)flow velocity(first moment); comparing both flow regimes (R-I and R-II), the discrepancies in the
√̅̅̅̅̅̅̅̅ mean velocity fields around the reference cylinder are significant (cf.
σv = 〈v2 〉 − dispersion of turbulent fluctuations (second moment); Fig. 8a.1 and 8a.2 for Vx , Fig. 9a.1 and 9a.2 for Vy ), while these are
noticeably smaller for the rough (cf. Fig. 8b.1 and 8b.2 for Vx and
〈v3 〉 Fig. 9b.1 and 9b.2 for Vy ) and finned ones (cf. Fig. 8c.1 and 8c.2 for Vx
Sv = − skewness coefficient (third moment);
σ 3v and Fig. 9c.1 and 9c.2 for Vy ). It is also visible from Figs. 8–11 that, for
each cylinder, the first two moments of the velocity field – streamwise
and transverse components of the mean velocity (Vx and Vy ) and
dispersion of turbulent fluctuations (σvx and σvy ) – overall grow in
magnitude when a transition from R-I to R-II occurs, as well as the time-
averaged vapor content 〈α〉 (Fig. 12).

Fig. 5. Spatial distributions of the excess coefficient of transverse velocity


fluctuations over and in the wake of the smooth (reference) cylinder in R-I
(Table 2) resulting from the statistical filtration of the ensemble of instanta­
neous velocity fields [28] for different sample volumes N = (a) 1000 and (b) Fig. 6. Spatial distributions of the time-averaged vapor fraction over and in the
5000 realizations. The cylinder mask is composed of several parts as follows. Its wake of the smooth (reference) cylinder in R-II (Table 2) obtained by the
longitudinal section coinciding with the PIV measurement plane is highlighted evaluation method of vapor phase occurrence developed by Pervunin et al. [52]
by pink contour with the diagonal hatching, while the green contour with the for different sample volumes N = (a) 1000 and (b) 5000 realizations. The black
inverse diagonal hatching corresponds to the frontal end surface of the cylinder diagonal crosses in image (a) display the sampling points where distributions of
adjacent to the test channel sidewall. The tilted pink lines under the test body velocity fluctuations are further analyzed. The color mask is the same as
depict the borders of its shadow in the measurement plane. in Fig. 5.

7
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

cylinders. A similar effect was revealed by Tian et al. [64] and Zhang
et al. [76], where such suppression of turbulence in the streamwise di­
rection along with shear stresses was attributed to the wall hydropho­
bicity. It is also worth noting that, in those two papers, the
measurements were taken within the boundary layers on surfaces with
various hydrophobic properties, unlike the turbulent wakes behind the
cylinders with different wall roughness as in this article.
With a transition from R-I to R-II, the amplitude of the higher-order
moments (Svx , Svy , Evx , Evy ) also grows in case of the reference cylinder
(Figs. 13a–16a), while the expansion rate of the wake region with non-
zero values of the skewness and excess coefficients, which is typical of a
non-Gaussian distribution of velocity fluctuations, increases consider­
ably. As for the rough and finned surfaces, these effects are almost ab­
sent: no significant changes are visible in the distributions of the higher-
order moments, either in magnitude or in the location and extension of
the region of non-zero values (Figs. 13b–16b and Figs. 13c–16c). Here,
we especially note that Svx and Svy remain individually similar for the
two modified cylinders, not only when the flow conditions are varied,
but also for the two distinctive levels of roughness which differ by nearly
two orders of magnitude. To explain this fact, the reader should recall
that turbulent vortices, which size falls within the inertial interval of a
turbulence energy spectrum, are characterized by homogeneous
Fig. 7. Histograms of the repeatability of vapor phase occurrence in R-II isotropic properties.
(Table 2) at three points (0.7; 0), (1.1; 0) and (1.5; 0) shown in Fig. 5a for The effects of asymmetry and spotting of turbulent fluctuations
illustrative purposes, where i is an index, ni is the number of i consecutive
expressed essentially by the coefficients of skewness and excess are
detections of the dispersed phase in the measured ensemble of instantaneous
manifested by integral vortices from the long-wave energy-containing
velocity fields.
range and Kolmogorov eddies from the viscous range of this spectrum (i.
e., large-scale and dissipative small-scale vortices of the order of the
This increase is, however, less pronounced for the rough and finned
Kolmogorov microscale). Given that Kolmogorov eddies quickly decay
cylinders compared to the reference one (e.g., cf. Figs. 10b, 10c, 11b and
(dissipate), the above-mentioned effects reflect the influence of rough­
11c against Figs. 10a and 11a). Moreover, in case of the smooth surface
ness mainly on large-scale vortex structures in the wake of the cylinder.
in contrast to the two others, the rise of the transverse component of
Thus, in case of the rough wall, an increase in the bulk flow velocity or,
velocity fluctuations significantly exceeds that of the streamwise one
which is equivalent, a decrease in the cavitation number has little effect
(nearly one and a half times). Thus, the anisotropy of turbulent fluctu­
on the structure of large-scale vortices (i.e., asymmetry and excess), and
ations enhances, which is not the case for the rough and finned
their skewness coefficient is almost independent of the scale of

Fig. 8. Spatial distributions of the streamwise component of the mean flow velocity around and in the wake of the (a) smooth, (b) rough and (c) finned cylinder in
the two flow regimes (Table 2). The color mask is the same as in Fig. 5.

8
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

Fig. 9. Spatial distributions of the transverse component of the mean flow velocity around and in the wake of the (a) smooth, (b) rough and (c) finned cylinder in the
two flow regimes (Table 2). The color mask is the same as in Fig. 5.

Fig. 10. Spatial distributions of the dispersion of streamwise velocity fluctuations around and in the wake of the (a) smooth, (b) rough and (c) finned cylinder in the
two flow regimes (Table 2). The color mask is the same as in Fig. 5.

roughness. surface hydrophobicity created by micro- or nano-textures pushes large-


The generation and evolution of these vortices is mainly determined scale vortices away from the wall, thereby reducing its influence on
by the presence, size and shape of the test body, as opposed to the them. Although the scale of roughness in the present study is different,
smooth cylinder. As noted by Tian et al. [64] and Zhang et al. [76], the the mechanism of its interaction with integral eddies and, therefore, its

9
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

Fig. 11. Spatial distributions of the dispersion of transverse velocity fluctuations around and in the wake of the (a) smooth, (b) rough and (c) finned cylinder in the
two flow regimes (Table 2). The color mask is the same as in Fig. 5.

influence on the higher-order moments of turbulent fluctuations is The distributions of the time-averaged vapor fraction shown in
obviously similar to that described by Tian et al. [64] and Zhang et al. Fig. 12 demonstrate the effect of noticeable reduction of the vapor
[76]. It is also visible from Fig. 13a and 13b that the fields of Svx for the concentration by roughness. It can be seen that, in regime R-I, the
smooth and rough surfaces are absolutely different in some region at a dispersed phase is only present on the surface of the polished cylinder
distance of about 2D downstream of the cylinder rear stagnation point in and in the mixing layers downstream (Fig. 12a), with no vapor found
both flow regimes (negative values in the reference case and positive behind the rough and finned ones (not shown). In R-II, cavitation be­
ones for the rough wall). The magnitude of Svy in a region of an comes further developed on the polished sample, where vapor already
approximate extension of 1D immediately behind the smooth cylinder is covers the entire field of the wake flow (Fig. 12b). In this regime, the
positive above and negative below the wake axis (Fig. 14a) in contrast to presence of the dispersed phase is also characteristic of both modified
both modified bodies, for which Svy is entirely positive across this region cylinders, with a significantly reduced vapor content (Fig. 12c and 12d).
(Fig. 14b and 14c). However, it is seen that the areas within both mixing layers, where the
The type of roughness affects the asymmetry of velocity fluctuations dispersed phase fraction takes its maximum values, are shifted down­
in both directions solely by changing the length of the indicated area, stream relative to the cylinder compared to the smooth sample in R-I
which differs by nearly 50 % (cf. Fig. 13b and 13c, Fig. 14b and 14c). In (Fig. 12a) because vapor must be chiefly concentrated in the cores of
this case, for both regimes (R-I and R-II), the distributions of the strong Karman vortices. This effect of cavitation mitigation is obviously
skewness coefficient of each velocity component in the wakes of the associated with increased slip on any of the rough walls (see Introduc­
modified cylinders almost do not change in contrast to the reference one tion). This also explains the downstream displacement of vortex struc­
(cf. Figs. 13b, 13c, 14b and 14c against Figs. 13a and 14a). This tures in the wake (cf. Fig. 12a, 12c and 12d).
discrepancy is evidently linked with the mechanism by which the small- It is observable from Fig. 8 that, for the polished sample, the region of
scale irregular roughness influences the skewness coefficient in the reverse flow is located directly behind the cylinder in R-I but it is absent
boundary layer, as discussed by Keirsbulck et al. [39] and Bhaganagar in R-II due to turbulization of the wake (Fig. 8a), whereas for the two
et al. [7]. This effect on Svx reflects a significant change in the properties rough bodies this region is visible in both regimes and significantly
of large-scale vortex structures in the boundary layer and, as a conse­ stretched downstream, with its maximum located at nearly 1.5D behind
quence, in the cylinder wake. The great influence of roughness on the the rear of the cylinder (Fig. 8b and 8c). It is important to note that the
excess coefficient in both directions can be clearly observed in Figs. 15 small-scale irregular wall morphology is more effective than the large-
and 16 in contrast to the skewness coefficient, for which this effect is scale regular texture in alleviating cavitation, such that the difference
chiefly manifested on its streamwise component (cf. Figs. 13 and 14). in the time-averaged vapor fraction reaches an order of magnitude.
This difference is especially distinctive in both regimes (R-I and R-II) for Besides, comparing the results with those from Churkin et al. [11] ob­
Evy in the wake, at a distance of 2D behind the cylinder rear edge (cf. tained for similar flow conditions and wall morphology, the small-scale
Fig. 16a and 16b), which is consistent with the results by Bhaganagar roughness is much more effective on the cylinder than on the NACA0015
et al. [7]. Similar to the asymmetry, the distributions of both compo­ hydrofoil due presumably to the significant surface curvature and faster
nents of the excess coefficient in the wakes of the modified test bodies separation of the boundary layer on the bluff body. Zaresharif et al. [74]
are practically identical, regardless of the flow conditions, in contrast to reported that the intensity of cavitation can be reduced with the aid of
the reference one (cf. Figs. 15b, 15c, 16b and 16c against Figs. 15a and some laser-made textures as compared to polished surfaces. This is also
16a). consistent with our measurements.

10
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

Fig. 12. Spatial distributions of the time-averaged vapor content and instantaneous images of cavitation structures at the moment of maximum cavitation intensity
around and in the wake of the (a) smooth, (b) rough and (c) finned cylinder in the two flow regimes (Table 2). The color mask is the same as in Fig. 5.

3.3. Distributions of fluctuating velocity roughness (Figs. 17c and 18c). According to Ilyushin et al. [31], such
a two-mode form of the histograms is associated with the process of
In addition to the above analysis of the flow turbulence structure, periodic advection of large-scale vortices in the wake of the cylinder,
below we consider histograms of velocity fluctuations in certain points with the peaks of the distributions corresponding to the utmost (oppo­
of the cylinder wake to delve deeper into the mechanism of interaction site) phases of these flow oscillations.
between turbulence and wall roughness. Figs. 17 and 18 disclose the Ilyushin et al. [31] also demonstrated that a two-peak histogram of
anisotropic nature of velocity fluctuations in the wake as noted above. the fluctuating velocity is typical of unsteady cloud cavitation over a
Moreover, the anisotropy manifests itself not only in the dispersion hydrofoil. The relatively high coherence of vapor cloud detachments
value, but also in the statistical structure of turbulent fluctuations (i.e., allowed the authors to collate certain phases of the quasi-periodic flow
difference in the appearance of the histograms for both velocity com­ dynamic with each of the peaks of the bimodal distribution. Coutier-
ponents). Particular attention should be drawn to their bimodality for Delgosha et al. [14] concluded that wall roughness in the range from
the transverse fluctuations (i.e., normal to the wall in the flow separa­ 100 to 400 μm caused disorganization of the growth/reduction cycle of
tion region) for the reference cylinder with negligible roughness. It can unsteady cloud cavitation, with this effect become more pronounced
be seen in Figs. 17a and 18a that, for the reference body, the distribu­ with increasing the level of irregularities. Apparently, such a disorga­
tions of the transverse fluctuating velocity vy have a bimodal shape (i.e., nization of the cloud (vortex) shedding process leads to a more chaotic
with two definite peaks) throughout the entire wake. The same quantity character of liquid motion in the wake of the rough and finned cylinders
in case of the cylinder with the small-scale roughness also tends to have and, as a result, to disruption of the phase coherence of the auto-
a bimodal appearance of its distribution, with this effect becoming more oscillation cycle of the sheet cavity length and cloud shedding, thus
distinctive when shifting downstream from the test body (Figs. 17b and causing a transformation of the distribution from bimodal into single-
18b). However, this tendency is practically invisible for the large-scale mode. Besides, highly transient flow conditions behind the test bodies

11
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

Fig. 13. Spatial distributions of the skewness coefficient of streamwise velocity fluctuations around and in the wake of the (a) smooth, (b) rough and (c) finned
cylinder in the two flow regimes (Table 2). The color mask is the same as in Fig. 5.

Fig. 14. Spatial distributions of the skewness coefficient of transverse velocity fluctuations around and in the wake of the (a) smooth, (b) rough and (c) finned
cylinder in the two flow regimes (Table 2). The color mask is the same as in Fig. 5.

in this case do not allow a vapor sheet to form and develop, disturbing it 250 μm, Ma et al. [45] showed that the larger the scale of irregularities
continuously, which is a possible reason of the above-mentioned sup­ on a cylinder surface is, the more effectively bistable states can be
pression of cavitation. Finally, considering roughness scales from 2 to inhibited.

12
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

Fig. 15. Spatial distributions of the excess coefficient of streamwise velocity fluctuations around and in the wake of the (a) smooth, (b) rough and (c) finned cylinder
in the two flow regimes (Table 2). The color mask is the same as in Fig. 5.

Fig. 16. Spatial distributions of the excess coefficient of transverse velocity fluctuations around and in the wake of the (a) smooth, (b) rough and (c) finned cylinder
in the two flow regimes (Table 2). The color mask is the same as in Fig. 5.

4. Conclusions Particle Image Velocimetry to measure velocity fields, along with the
procedure of statistical filtration of velocity vectors and the method of
Using high-speed visualization to observe cavitation structures and vapor phase detection both developed and described in our previous

13
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

Fig. 17. Histograms of both components of the fluctuating velocity in the wake of the (a) smooth, (b) rough and (c) finned cylinder at distances of 1D (see point (1.5;
0) in Figs. 8–16), 2D (2.5; 0) and 3D (3.5; 0) downstream of its rear stagnation point in R-I according to Table 2. The color legend is the same for all the histograms.

publications, we performed an experimental study on cavitation sup­ 3. Irrespective of the type of surface morphology, the influence of
pression and transformation of turbulence structure in the transverse roughness on the mechanism of formation of large-scale vortices and
flow around a circular cylinder through surface modification. For this, their characteristics is weakened. However, it leads to overall
we used three identical test bodies with different wall morphology, disorganization of liquid motion in the wake, thus making local flow
namely smooth one (reference), micro-scale irregularities (rough) and conditions highly unsteady. The larger the scale of irregularities is,
large-scale (of the order of a millimeter) regular texture (finned). the more chaotic this process becomes.
Although the materials, which the test bodies were made of or covered
with, were demonstrated to be essentially hydrophilic, the rough sample These findings also provide useful information for design engineers
exhibited stable and pronounced hydrophobic properties thanks to the on promising materials, advanced coating and surface treatment tech­
specific morphology of its surface. nologies and effective roughness morphologies that can be hypotheti­
The conducted in-depth analysis of the visual data and measured cally used in practice to improve the composition of underwater
turbulence characteristics of the flow velocity, including the higher- components of hydraulic units and propulsion systems. Future research
order moments of turbulent fluctuations (i.e., the coefficients of skew­ in this area should be focused on the influence of isolated irregularities
ness and excess) and time-averaged vapor content around and in the of various wall morphologies and scales featuring different wetting
wake of each of the cylinders in the two typical regimes of the cavitating properties on the local structure of velocity field, nucleation and evo­
flow allowed us to conclude the following: lution of individual vapor bubbles, their aggregation into a cavity and its
subsequent detachment. This evidently requires highly spatio-
1. Elevated surface roughness promotes excellent cavitation suppres­ temporally resolved measurements, distinguishing the continuous
sion. For the considered wall morphologies, the smaller the scale of (liquid) and dispersed (vapor) phases, to gain further insights into the
irregularities is, the more pronounced this effect occurs to be, indi­ mechanism of cavitating flow control by the passive means.
cating the high efficiency of this passive control.
2. Both types of roughness noticeably affect the turbulence structure of CRediT authorship contribution statement
the wake flow, including the mean velocity, dispersion and higher-
order moments of turbulent fluctuations. A change in the flow Mikhail Yu. Nichik: Visualization, Validation, Investigation, Data
regime for the rough and finned cylinders has almost no effect on the curation. Boris B. Ilyushin: Formal analysis, Data curation, Methodol­
amplitude of the higher-order moments, regardless of the roughness ogy, Software, Writing – original draft, Writing - review & editing.
scale, in contrast to the reference one. Ebrahim Kadivar: Supervision, Resources, Project administration,

14
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

Fig. 18. Histograms of both components of the fluctuating velocity in the wake of the (a) smooth, (b) rough and (c) finned cylinder at distances of 1D (see point (1.5;
0) in Figs. 8–16), 2D (2.5; 0) and 3D (3.5; 0) downstream of its rear stagnation point in R-II according to Table 2. The color legend is the same for all the histograms.

Conceptualization. Ould el Moctar: Funding acquisition, Conceptuali­ [2] R.J. Adrian, Hairpin vortex organization in wall turbulence, Phys. Fluids 19 (4)
(2007), https://doi.org/10.1063/1.2717527 pp. (041301)–16.
zation. Konstantin S. Pervunin: Formal analysis, Supervision, Valida­
[3] Arndt, R.E.A., Ippen, A.T., 1967. Cavitation near surfaces of distributed roughness.
tion, Writing – original draft, Writing - review & editing. Massachusetts Institute of Technology, Cambridge. Technical report No. 104, p.
164.
[4] R.E.A. Arndt, A.T. Ippen, Rough surface effects on cavitation inception, ASME J.
Basic Eng. 90 (2) (1968) 249–261, https://doi.org/10.1115/1.3605086.
Declaration of competing interest
[5] K. Autumn, Y.A. Liang, S.T. Hsieh, W. Zesch, W.P. Chan, T.W. Kenny, R. Fearing, R.
J. Full, Adhesive force of a single gecko foot-hair, Nature 405 (2000) 681–685,
The authors declare that they have no known competing financial https://doi.org/10.1038/35015073.
[6] W. Barthlott, C. Neinhuis, Purity of the sacred lotus, or escape from contamination
interests or personal relationships that could have appeared to influence
in biological surfaces, Planta 202 (1) (1997) 1–8, https://doi.org/10.1007/
the work reported in this paper. s004250050096.
[7] K. Bhaganagar, J. Kim, G. Coleman, Effect of roughness on wall-bounded
turbulence, Flow Turbul. Combust. 72 (2004) 463–492, https://doi.org/10.1023/
Acknowledgements B:APPL.0000044407.34121.64.
[8] H.M. Blackburn, W.H. Melbourne, The effect of free-stream turbulence on sectional
The manufacturing of the test bodies was funded by DFG (Project No. lift forces on a circular cylinder, J. Fluid Mech. 306 (1996) 267–292, https://doi.
org/10.1017/S0022112096001309.
469042952), the experiment on the cavitating flow around the cylinders [9] A.V. Boiko, G.R. Grek, A.V. Dovgal, V.V. Kozlov, The origin of turbulence in near-
with different surface properties was carried out with a financial support wall flows, 10.1007/978-3-662-04765-1, Springer, Berlin, Heidelberg, 2002,
from the Russian Science Foundation (Project No. 19-79-30075-P) and p. 278.
[10] P.A. Brandner, G.J. Walker, P.N. Niekamp, B. Anderson, An experimental
the analysis of the turbulence flow structure was conducted under a state
investigation of cloud cavitation about a sphere, J. Fluid Mech. 656 (2010)
contract with IT SB RAS. The authors express their deep gratitude to Mr. 147–176, https://doi.org/10.1017/S0022112010001072.
Mikhail Timoshevskiy for his valuable assistance in preparing the test [11] S.A. Churkin, K.S. Pervunin, A.Y. Kravtsova, D.M. Markovich, K. Hanjalic,
Cavitation on NACA0015 hydrofoils with different wall roughness: High-speed
bodies for the experiment and Dr. Vyacheslav Cheverda for his kind help
visualization of the surface texture effects, J. Vis. 19 (4) (2016) 587–590, https://
in measuring the contact angles of the materials. doi.org/10.1007/s12650-016-0355-9.
[12] K.-S. Choi, N. Fujisawa, Possibility of drag reduction using d-type roughness, Appl.
Sci. Res. 50 (1993) 315–324, https://doi.org/10.1007/BF00850564.
References [13] C. Cottin-Bizonne, B. Cross, A. Steinberger, E. Charlaix, Boundary slip on smooth
hydrophobic surfaces: Intrinsic effects and possible artifacts, pp. (056102)–4, Phys.
[1] E. Achenbach, Influence of surface roughness on the cross-flow around a circular Rev. Lett. 94 (5) (2005), https://doi.org/10.1103/PhysRevLett.94.056102.
cylinder, J. Fluid Mech. 46 (2) (1971) 321–335, https://doi.org/10.1017/
S0022112071000569.

15
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

[14] O. Coutier-Delgosha, J.-F. Devillers, M. Leriche, T. Pichon, Effect of wall roughness [42] G.V. Kuznetsov, A.G. Islamova, E.G. Orlova, A.S. Ivashutenko, I.I. Shanenkov, I.
on the dynamics of unsteady cavitation, ASME J. Fluids Eng. 127 (4) (2005) Y. Zykov, D.V. Feoktistov, Influence of roughness on polar and dispersed
726–733, https://doi.org/10.1115/1.1949637. components of surface free energy and wettability properties of copper and steel
[15] A. Danlos, J.-E. Méhal, F. Ravelet, O. Coutier-Delgosha, F. Bakir, Study of the surfaces, pp. (127518)–12, Surf. Coat. Technol. 422 (2021), https://doi.org/
cavitating instability on a grooved Venturi profile, pp. (101302)–10, ASME J. 10.1016/j.surfcoat.2021.127518.
Fluids Eng. 136 (10) (2014), https://doi.org/10.1115/1.4027472. [43] A.T. Leger, S.L. Ceccio, Examination of the flow near the leading edge of attached
[16] J.A. Dean, Lange’s Handbook of Chemistry, 15th ed., McGraw-Hill Inc., New York, cavitation. Part 1. Detachment of two-dimensional and axisymmetric cavities,
1999. J. Fluid Mech. 376 (1998) 61–90, https://doi.org/10.1017/s0022112098002766.
[17] L. Dellieu, M. Sarrazin, P. Simonis, O. Deparis, J.P. Vigneron, A two-in-one [44] Y. Li, H. Chen, J. Wang, D. Chen, Effect of grooves on cavitation around the body of
superhydrophobic and anti-reflective nanodevice in the grey cicada Cicada orni revolution. ASME, pp. (011301)–7, J. Fluids Eng. 132 (1) (2010), https://doi.org/
(Hemiptera), pp. (024701)–8, J. Appl. Phys. 116 (2) (2014), https://doi.org/ 10.1115/1.4000648.
10.1063/1.4889849. [45] W. Ma, Q. Liu, J.H.G. Macdonald, X. Yan, Y. Zheng, The effect of surface roughness
[18] X. Deng, L. Mammen, H.-J. Butt, V. Doris, Candle soot as a template for a on aerodynamic forces and vibrations for a circular cylinder in the critical
transparent robust superamphiphobic coating, Science 335 (6064) (2011) 67–70, Reynolds number range, J. Wind Eng. Ind. Aerodyn. 187 (2019) 61–72, https://
https://doi.org/10.1126/science.1207115. doi.org/10.1016/j.jweia.2019.01.011.
[19] M. Dular, B. Bachert, B. Stoffel, B. Širok, Relationship between cavitation [46] A.Y. Malkin, S.A. Patlazhan, V.G. Kulichikhin, Physicochemical phenomena
structures and cavitation damage, Wear 257 (11) (2004) 1176–1184, https://doi. leading to slip of a fluid along a solid surface, Russ. Chem. Rev. 88 (3) (2019)
org/10.1016/j.wear.2004.08.004. 319–349, https://doi.org/10.1070/RCR4849.
[20] J.P. Franc, J.M. Michel, Unsteady attached cavitation on an oscillating hydrofoil, [47] K. Manoharan, S. Bhattacharya, Superhydrophobic surfaces review: Functional
J. Fluid Mech. 193 (1988) 171–189, https://doi.org/10.1017/ application, fabrication techniques and limitations, J. Micromanuf. 2 (1) (2019)
s0022112088002101. 59–78, https://doi.org/10.1177/2516598419836345.
[21] J.-P. Franc, J.-M. Michel, Fundamentals of cavitation. In: Fluid Mechanics and its [48] I.V. Marchuk, V.V. Cheverda, P.A. Strizhak, O.A. Kabov, Determination of surface
Applications 76 (2004) 328. ISBN 1-4020-2232-8. https://doi.org/10.1007/1- tension and contact angle by the axisymmetric bubble and droplet shape analysis,
4020-2233-6. Thermophys. Aeromech. 22 (3) (2015) 297–303, https://doi.org/10.1134/
[22] X. Gao, L. Jiang, Water-repellent legs of water striders, Nature 432 (2004) 36, S0869864315030038.
https://doi.org/10.1038/432036a. [49] M.Y. Nichik, M.V. Timoshevskiy, K.S. Pervunin, Effect of an end-clearance width
[23] M. Ghoohestani, S. Rezaee, E. Kadivar, M.A. Esmaeilbeig, Reactive-dynamic on the gap cavitation structure: Experiments on a wall-bounded axis-equipped
characteristics of a nanobubble collapse near a solid boundary using molecular hydrofoil, pp. (111387)–19, Ocean Eng. 254 (2022), https://doi.org/10.1016/j.
dynamic simulation, pp. (022003)–13, Phys. Fluids 35 (2) (2023), https://doi.org/ oceaneng.2022.111387.
10.1063/5.0139169. [50] A.R. Parker, C.R. Lawrence, Water capture by a desert beetle, Nature 414 (2001)
[24] M. Ghoohestani, S. Rezaee, E. Kadivar, O. el Moctar, Thermodynamic effects on 33–34, https://doi.org/10.1038/35102108.
nanobubble’s collapse-induced erosion using molecular dynamic simulation, pp. [51] R.F. Patella, T. Choffat, J.-L. Reboud, A. Archer, Mass loss simulation in cavitation
(073319)–11, Phys. Fluids 35 (7) (2023), https://doi.org/10.1063/5.0154822. erosion: Fatigue criterion approach, Wear 300 (1–2) (2013) 205–215, https://doi.
[25] K.B. Golovin, J.W. Gose, M. Perlin, S.L. Ceccio, A. Tuteja, Bioinspired surfaces for org/10.1016/j.wear.2013.01.118.
turbulent drag reduction, pp. (20160189)–20, Philos. Trans. R. Soc. A Math. Phys. [52] K.S. Pervunin, M.V. Timoshevskiy, B.B. Ilyushin, Distribution of probability of the
Eng. Sci. 374 (2073) (2016), https://doi.org/10.1098/rsta.2016.0189. vapor phase occurrence in a cavitating flow based on the concentration of PIV
[26] O. Güven, C. Farell, V.C. Patel, Surface-roughness effects on the mean flow past tracers in liquid, pp. (247)–12, Exp. Fluids 62 (12) (2021), https://doi.org/
circular cylinders, J. Fluid Mech. 98 (4) (1980) 673–701, https://doi.org/10.1017/ 10.1007/s00348-021-03344-y.
S0022112080000341. [53] M.R. Raupach, R.A. Antonia, S. Rajagopalan, Rough-wall turbulent boundary
[27] C. Haosheng, L. Yongjian, C. Darong, W. Jiadao, Experimental and numerical layers, ASME Appl. Mech. Rev. 44 (1) (1991) 1–25, https://doi.org/10.1115/
investigations on development of cavitation erosion pits on solid surface, Tribol. 1.3119492.
Lett. 26 (2007) 153–159, https://doi.org/10.1007/s11249-006-9188-3. [54] G.E. Reisman, Y.-C. Wang, C.E. Brennen, Observations of shock waves in cloud
[28] O. Heinz, B. Ilyushin, D. Markovich, Application of a PDF method for the statistical cavitation, J. Fluid Mech. 355 (1998) 255–283, https://doi.org/10.1017/
processing of experimental data, Int. J. Heat Fluid Flow 25 (5) (2004) 864–874, S0022112097007830.
https://doi.org/10.1016/j.ijheatfluidflow.2004.05.009. [55] J.L.D. Ribeiro, Effects of surface roughness on the two-dimensional flow past
[29] J.W. Holl, The inception of cavitation on isolated surface irregularities, ASME J. circular cylinders I: Mean forces and pressures, J. Wind Eng. Ind. Aerodyn. 37 (3)
Basic Eng. 82 (1) (1960) 169–183, https://doi.org/10.1115/1.3662508. (1991) 299–309, https://doi.org/10.1016/0167-6105(91)90014-N.
[30] Y. Iga, J. Okajima, Y. Yamagichi, H. Sasaki, Y. Ito, Thermodynamic suppression [56] M. Sadri, E. Kadivar, Numerical investigation of the cavitating flow and the
effect of cavitation arising in a hydrofoil in 140 ◦ C hot water, pp. (011207)–9, cavitation-induced noise around one and two circular cylinders, pp. (114178)–16,
ASME J. Fluids Eng. 145 (1) (2023), https://doi.org/10.1115/1.4055600. Ocean Eng. 277 (2023), https://doi.org/10.1016/j.oceaneng.2023.114178.
[31] B.B. Ilyushin, M.V. Timoshevskiy, K.S. Pervunin, Vapor concentration and bimodal [57] H.K.G. Schlichting, Boundary-layer theory, 9 ed., Springer Berlin, Heidelberg,
distributions of turbulent fluctuations in cavitating flow around a hydrofoil, pp. 2016, p. 833.
(109197)–12, Int. J. Heat Fluid Flow 103 (2023), https://doi.org/10.1016/j. [58] A.V. Sentyabov, M.V. Timoshevskiy, K.S. Pervunin, Gap cavitation in the end
ijheatfluidflow.2023.109197. clearance of a guide vane of a hydroturbine: Numerical and experimental
[32] H. Jia, R. Xie, Y. Zhou, Experimental investigation of the supercavitation and investigation, J. Eng. Thermophys. 28 (1) (2019) 67–83, https://doi.org/10.1134/
hydrodynamic characteristics of high-speed projectiles with hydrophobic and S1810232819010065.
hydrophilic coatings, pp. (363)–15, MDPI Fluids 7 (12) (2022), https://doi.org/ [59] S.K. Sethi, G. Manik, Recent progress in super hydrophobic/hydrophilic self-
10.3390/fluids7120363. cleaning surfaces for various industrial applications: A review, Polym.-Plast.
[33] J. Jiménez, Turbulent flows over rough walls, Annu. Rev. Fluid Mech. 36 (1) Technol. Eng. 57 (18) (2018) 1932–1952, https://doi.org/10.1080/
(2004) 173–196, https://doi.org/10.1146/annurev.fluid.36.050802.122103. 03602559.2018.1447128.
[34] M.A. Jouybari, J. Yuan, G.J. Brereton, M.S. Murillo, Data-driven prediction of the [60] M.H. Shim, J. Kim, C.H. Park, The effects of surface energy and roughness on the
equivalent sand-grain height in rough-wall turbulent flows, pp. (A8)–23, J. Fluid hydrophobicity of woven fabrics, Text. Res. J. 84 (12) (2014) 1268–1278, https://
Mech. 912 (2021), https://doi.org/10.1017/jfm.2020.1085. doi.org/10.1177/0040517513495945.
[35] Y.C. Jung, B. Bhushan, Contact angle, adhesion and friction properties of micro- [61] L. Solovyeva, N. Zubkov, B. Lisowsky, A. Elmoursi, Novel electrical joints using
and nanopatterned polymers for superhydrophobicity, Nanotechnology 17 (19) deformation machining technology – Part II: Experimental verification, IEEE Trans.
(2006) 4970–4980, https://doi.org/10.1088/0957-4484/17/19/033. Compon. Packag. Manuf. Technol. 2 (10) (2012) 1718–1722, https://doi.org/
[36] E. Kadivar, Experimental and Numerical Investigations of Cavitation Control Using 10.1109/TCPMT.2012.2199755.
Cavitating-Bubble Generators, University of Duisburg-Essen, 2020. PhD thesis. [62] Q. Sun, The hydrophobic effects: Our current understanding, MDPI Molecules 27
[37] E. Kadivar, T. Ochiai, Y. Iga, O. el Moctar, An experimental investigation of (20) (2022), https://doi.org/10.3390/molecules27207009 pp. (7009)–27.
transient cavitation control on a hydrofoil using hemispherical vortex generators, [63] A. Theophanatos, J. Wolfram, Hydrodynamic loading on macro-roughened
J. Hydrodyn. 33 (2021) 1139–1147, https://doi.org/10.1007/s42241-021-0097-6. cylinders of various aspect ratios, ASME J. Offshore Mech. Arctic Eng. 111 (3)
[38] M. Kadivar, D. Tormey, G. McGranaghan, A review on turbulent flow over rough (1989) 214–222, https://doi.org/10.1115/1.3257150.
surfaces: Fundamentals and theories, pp. (100077)–34, Internat. J. Thermofluids [64] H. Tian, J. Zhang, N. Jiang, Z. Yao, Effect of hierarchical structured
10 (2021), https://doi.org/10.1016/j.ijft.2021.100077. superhydrophobic surfaces on coherent structures in turbulent channel flow, Exp.
[39] L. Keirsbulck, L. Labraga, A. Mazouz, C. Tournier, Surface roughness effects on Therm Fluid Sci. 69 (2015) 27–37, https://doi.org/10.1016/j.
turbulent boundary layer structures, ASME J. Fluids Eng. 124 (1) (2002) 127–135, expthermflusci.2015.07.018.
https://doi.org/10.1115/1.1445141. [65] M.V. Timoshevskiy, B.B. Ilyushin, K.S. Pervunin, Statistical structure of the velocity
[40] M. Kosuda, Y. Kubota, M. Yokoyama, O. Mochizuki, Introduction of multiscaled field in cavitating flow around a 2D hydrofoil, pp. (108646)–11, Int. J. Heat Fluid
longitudinal vortices by fractal-patterned surface roughness, J. Flow Control Flow 85 (2020), https://doi.org/10.1016/j.ijheatfluidflow.2020.108646.
Measure. Visualizat. 7 (2) (2019) 120–132, https://doi.org/10.4236/ [66] V. Ulianitsky, A. Shtertser, S. Zlobin, I. Smurov, Computer-controlled detonation
jfcmv.2019.72010. spraying: From process fundamentals toward advanced applications, J. Therm.
[41] A.Y. Kravtsova, D.M. Markovich, K.S. Pervunin, M.V. Timoshevskiy, K. Hanjalić, Spray Technol. 20 (4) (2011) 791–801, https://doi.org/10.1007/s11666-011-
High-speed visualization and PIV measurements of cavitating flows around a semi- 9649-6.
circular leading-edge flat plate and NACA0015 hydrofoil, Int. J. Multiph. Flow 60 [67] M. van Rijsbergen, A review of sheet cavitation inception mechanisms. Open
(2014) 119–134, https://doi.org/10.1016/j.ijmultiphaseflow.2013.12.004. Archives of 16th International Symposium on Transport Phenomena and Dynamics

16
M. Yu. Nichik et al. Ultrasonics Sonochemistry 106 (2024) 106875

of Rotating Machinery (ISROMAC 2016), April 10–15, 2016. Honolulu, Hawaii, [73] H. Ye, L. Zhu, W. Li, H. Liu, H. Chen, Simple spray deposition of a water-based
US, Article No. hal-01890067. https://hal.science/hal-01890067. superhydrophobic coating with high stability for flexible applications, J. Mater.
[68] J.A. Venning, B.W. Pearce, P.A. Brandner, Nucleation effects on cloud cavitation Chem. A 5 (20) (2017) 9882–9890, https://doi.org/10.1039/C7TA02118F.
about a hydrofoil, pp. (A1)–26, J. Fluid Mech. 947 (2022), https://doi.org/ [74] M. Zaresharif, F. Ravelet, D.J. Kinahan, Y.M.C. Delaure, Cavitation control using
10.1017/jfm.2022.535. passive flow control techniques, pp. (121301)–35, Phys. Fluids 33 (12) (2021),
[69] Z. Wang, H. Cheng, B. Ji, Euler-Lagrange study of cavitating turbulent flow around https://doi.org/10.1063/5.0071781.
a hydrofoil, pp. (112108)–18, Phys. Fluids 33 (11) (2021), https://doi.org/ [75] M.M. Zdravkovich, Flow around Circular Cylinders. Volume I: Fundamentals,
10.1063/5.0070312. Oxford Science Publications, Oxford, UK, 1997, p. 672. ISBN 9780198563969.
[70] Z. Wang, H. Cheng, B. Ji, X. Peng, Numerical investigation of inner structure and its [76] J. Zhang, H. Tian, Z. Yao, P. Hao, N. Jiang, Mechanisms of drag reduction of
formation mechanism of cloud cavitating flow, pp. (104484)–19, Int. J. Multiph. superhydrophobic surfaces in a turbulent boundary layer flow, pp. (179)–13, Exp.
Flow 165 (2023), https://doi.org/10.1016/j.ijmultiphaseflow.2023.104484. Fluids 56 (9) (2015), https://doi.org/10.1007/s00348-015-2047-y.
[71] G.S. West, C.J. Apelt, The effects of tunnel blockage and aspect ratio on the mean [77] B. Zhou, X. Wang, W.M. Gho, S.K. Tan, Force and flow characteristics of a circular
flow past a circular cylinder with Reynolds numbers between 104 and 105, J. Fluid cylinder with uniform surface roughness at subcritical Reynolds numbers, Appl.
Mech. 114 (1982) 361–377, https://doi.org/10.1017/S0022112082000202. Ocean Res. 49 (2015) 20–26, https://doi.org/10.1016/j.apor.2014.06.002.
[72] F. Xia, L. Jiang, Bio-inspired, smart, multiscale interfacial materials, Adv. Mater. 20 [78] N.N. Zubkov, Multitool deformation and cutting in applying fins to heat-exchanger
(15) (2008) 2842–2858, https://doi.org/10.1002/adma.200800836. pipe, Russ. Eng. Res. 35 (11) (2015) 859–863, https://doi.org/10.3103/
S1068798X15110209.

17

You might also like