This study aims to experimentally investigate the vertical parallel water entry of two identical spheres (in geometry and material) with different surface wettability (hydrophilic or hydrophobic) pairings. The spheres simultaneously impact the water surface with velocities ranging from 1.71 to 4.32 m s 1. The corresponding ranges of the impact Froude, Weber, and Reynolds numbers are 3.87 9.75, 816 5167, and 38.5 × 10 3 to 96.8 × 10 3, respectively. The spheres' lateral distances vary from 1.0 to 5.0 times the diameter. A high-speed photography system and image processing technique analyze the event dynamics, focusing on air-entrainment cavity behavior (shapes, closure, shedding), water flow features (Worthington jets, splashes), and sphere kinetics. Results for hydrophobic/hydrophobic cases show that even at the maximum lateral distance, a slightly asymmetric cavity forms, but deep-seal pinching occurs at a single point, similar to a single water entry scenario. As the lateral distance decreases, the spheres significantly influence each other's behavior, leading to the formation of a highly asymmetric air cavity and an oblique Worthington jet. In the case of a hydrophobic/hydrophilic pairing, vortices generated behind the hydrophilic sphere influence the air cavity development of the hydrophobic sphere. This can cause a secondary pinch-off, especially at low lateral distances. This effect becomes more pronounced at higher impact velocities. Additionally, at higher impact velocities and minimum lateral distance (direct contact between the spheres), a smaller cavity detaches from the hydrophobic sphere's cavity, attaches to the hydrophilic sphere, and moves with it. These different regimes result in varying descent velocities for the spheres.

When a solid object impacts an undisturbed, free water surface and penetrates the water, it initiates a complex, nonlinear, and transient flow event involving the interaction between the solid, liquid, and gas phases. This phenomenon, commonly referred to as “water entry” in classical mechanics, encompasses a wide range of dynamic behaviors and events. Among these are air cavity formation, evolution, and closure, as well as splash crown and dome formation, water jet generation, and the occurrence of cavity ripples and shedding. These phenomena can significantly impact the dynamics of the object's movement and induce changes in its kinematic behavior.1–8 For over a century, engineers and researchers have dedicated significant efforts to studying the topic of water entry, driven by its broad spectrum of applications. Some notable applications include the following:

  1. Industrial processes: Water entry plays a crucial role in spraying, painting, printing, fluid atomization, coatings, and surface cleaning applications.9–13 

  2. Naval engineering: Water entry phenomena are significant in areas like ship launching, slamming (impact of a ship hull with waves), and the development of naval weapons.13–21 

  3. Aerospace and aeronautical engineering: Water entry studies are relevant to seaplane and spacecraft landing, air-to-water weapons, and they offer insight into the behavior of bombs during water impact.22–26 

  4. Sports engineering: Water entry research contributes to the technical analysis of oars in rowing and athletes in diving.12,26,27

  5. Observations in nature: Water entry phenomena have been observed in various natural scenarios, such as lizards that can walk on water surfaces, seabirds that dive into the water, and beetles that walk upside down underneath the water surface.1,25,28–31

With the emergence and expansion of high-speed camera technology, the study and investigation of water entry phenomena have entered a new phase, enabling the examination of objects with diverse geometries and under distinctive conditions. This technological advancement has significantly enhanced the recording, accuracy, and analysis of related phenomena.32 It is noteworthy that a majority of the studies presented and reported thus far have primarily focused on the water entry of spherical bodies,4,9,13,16,17,25,33–36 since such projectiles are capable of generating and exhibiting a wide range of interesting phenomena during water entry, including various main cavity closure regimes. Furthermore, their axisymmetric geometry allows for the formation of nearly axisymmetric cavities, jets, and splashes. As a result, conducting tests with a single high-speed camera is generally sufficient. Additionally, spherical bodies tend to experience minimal rotation and deviation in their trajectory. Nevertheless, several researchers have also focused on investigating water entry phenomena involving objects with non-spherical geometries. Examples of such objects include conical and wedge-shaped projectiles,37–43 flat plates,1,14,44–48 vertical cylindrical and slender bodies,8,32,49–52 bodies with asymmetric nose shapes,53 and thick and thin hollow objects.54–58 Moreover, studies have explored the water entry behavior of objects with more complex geometries, such as 3D-printed birds and teardrop-shaped objects.31,59,60 It is important to note that apart from geometric shapes, researchers are also interested in various other factors. These include the material and surface properties of the projectile, the physical properties of the fluid, the angle and rotation rate of the projectiles when they enter the water, the free-water surface deformations due to forced entry and exit of bodies, and even scenarios involving microgravity.2–4,8,12,26,61–67

A review of the existing literature highlights that the field of water entry research has primarily concentrated on examining single projectile impacts on the water surface. However, the investigation of parallel (side-by-side) or tandem (consecutive) water entry involving multiple objects remains relatively unexplored. This topic can be regarded as a highly unresolved research subject within the field. It is important to note that this problem differs from the extensively studied scenario of fluid passing over multiple bodies, which typically focuses on one-phase fluid flow phenomena.68–74 The parallel or simultaneous entry of multiple projectiles can give rise to interactions that significantly influence the development of air cavities, splash curtains, and jets, which in turn can impact the movement and dynamics of the objects involved. The phenomenon of parallel water entry has diverse applications, including synchronized diving in Olympic competitions, parallel diving behaviors of seabirds, side-by-side entry of airborne projectiles into the water, water model experiments simulating molten iron bath desulphurization processes to predict penetration depth and residence time of agents, inkjet printing technology, simultaneous entry and exit of adjacent oars in rowing, extreme waves affecting offshore wind turbine platforms and offshore vessels or structures installed side-by-side, and the consideration of loads on catamarans' twin hulls, among others.

As mentioned, the number of published studies focusing on multiple-object water entry is limited, but the topic has received quite some attention in recent years. In 2010, Wang and Wang75 conducted a numerical analysis on the parallel synchronous water entry of cylinders to investigate the impact of various parameters, such as cylinder radius, the distance between cylinders, and their velocities, on the hydrodynamic forces. In 2014, Ern and Brosse76 conducted a study examining the interaction of two identical disks falling side-by-side into salty water, varying the aspect ratios. They observed that when the lateral distance between the disks' centers was less than 4.5 diameters, the projectiles repelled each other and that the repulsion kinematics was also dependent on the aspect ratio. Yousefnezhad and Zeraatgar77 utilized the boundary element method (potential theory) to analyze the parallel water entry of two wedges at a constant vertical velocity. They proposed a regression formula for determining the maximum pressure coefficient on the projectile surface for different wedge angles. In 2018, Yu78 investigated the cavity evolution and kinetics of two bodies entering the water in tandem, employing both numerical and experimental techniques. Jiaxing et al.79 conducted an empirical study on the water entry process of two parallel cylinders. They compared the cavity characteristics of single and double cylinders at various Froude numbers. Their findings revealed that the entire cavity exhibited good mirror symmetry, whereas each cylinder's cavity displayed significant asymmetry. Furthermore, their results demonstrated that for low-impact Froude numbers ( Fr I), the cavity closure followed a characteristic pinching-off pattern, with the pinching point moving upward as Fr I increased. When Fr I reached a critical value, the closure transitioned to a surface-seal pattern, and the closure point moved downward with increasing Fr I.

Wang and Lyu7 conducted an experimental analysis of the cavity dynamics and kinetics of two spheres impacting the water surface side-by-side. The surfaces of the spheres had a static contact angle ranging from 80 ° ± 10 °, indicating a hydrophilic nature. The impact velocity chosen for the experiments was above 14 m s 1, which exceeds the threshold velocity, above which cavity formation occurs for hydrophilic spheres. They investigated the influence of time intervals between releases and the lateral distances on the water entry process. Their findings revealed that at a lateral distance between the spheres of more than 5.5 diameters, both the cavity characteristics and kinetics of the spheres resembled those observed in single water entry scenarios. However, when considering the combined effect of release time intervals and lateral distances, they observed intriguing phenomena, such as the expansion and contraction of the walls of two adjacent cavities, the formation of an M-shaped splash, and more. For moderate time intervals and lateral distances of less than 1.5 diameters, one sphere was horizontally attracted to the other, resulting in a successive water entry pattern. In contrast, in cases of synchronous water entry, repulsion between the two spheres was observed. These phenomena were explained by the relative strength of expulsion and pressure forces acting on the spheres.

In 2020, Yun et al.80 conducted an experimental study to investigate the hydrodynamic features associated with oblique water entry of two tandem spheres. Their research revealed the formation of a suction region within the cavity created by the front sphere. This suction region resulted in the acceleration of the rear sphere, and in some cases, even led to a collision between the two spheres. In a separate experimental study, Yun et al.81 investigated vertical tandem water entry of two spheres. They observed that when the rear sphere impacted the upward water jet generated by the front sphere, a significant amount of kinematic energy was transferred to the flow. As a result, a considerably larger cavity formed behind the rear sphere compared to the case of a single-sphere water entry.

Rabbi et al.82 conducted an experimental study where they investigated the tandem water entry of two spheres. Their primary focus was to measure the forces exerted on the trailing sphere. They observed that a reduction in impact deceleration occurred for the rear sphere, which was influenced by the cavity dynamics generated by the front sphere and the relative timing of the trailing sphere's impact. To categorize the phenomena arising from these combined effects, they introduced a non-dimensional number called Matryoshka number. Four main regimes were identified: on-cavity, inside-cavity, on/inside-bowl, and on-Worthington jet modes. The results revealed that in the first, second, and fourth regimes, the impact force on the trailing sphere could decrease by up to 78 %. However, in the third regime, a surprising increase in over 400 % in the force was observed in the worst-case scenario.

Rabbi83 experimentally examined the simultaneous side-by-side water entry of hydrophobic spheres of the same size. He observed that the air cavities formed by the spheres are not influenced when the lateral center-to-center distance ( Δ x c) is higher than 4.0 D s (here, D s is the sphere's diameter). In this study, for Δ x c < 4.0 D s, two bowl-shaped asymmetric cavities were observed, which prevented a clean pinch-off; instead, two long air tubes formed (string shedding). The pinching time of both cavities monotonically decreased with increasing sphere distance approaching t pinch = 2.06 D s / 2 g, i.e., the relation for predicting the pinch-off time of single sphere water entry. Rabbi also proposed a potential flow model to explain experimental observations (e.g., the dynamic interaction between the cavities' walls). Furthermore, he showed that as Fr I increases, the pinching times decrease and confirmed that two tilted Worthington jets are created after pinch off, which was most noticeable for smaller Δ x c. For very small lateral distances, a combined upper cavity was observed, resulting in the formation of a single chaotic and merged Worthington jet.

In 2021, Lyu et al.84 conducted experimental investigations to examine the impact of the time interval between two tandem spheres entering the water on cavity evolution and the motion characteristics of the spheres. Their findings revealed that the time interval between the spheres influences the pattern of cavity closure for the rear sphere. Lu et al.85 examined high-speed parallel projectile's water entry (from 280 to 400 m s 1) numerically. At such impact velocities, their results show that the cavity distribution is nonuniform due to the strong interference between the projectiles. In 2022, Lu et al.86 numerically investigated the asynchronous parallel water entry of two slender cylindrical projectiles. They reported that, at first, the air entrainment cavity of the first projectile is squeezed, and the cavity of the second one is expanded, and eventually, both cavities merged and gradually stretched. In 2022, Lyu et al.87 experimentally and numerically analyzed the different modes of water entry for the trailing sphere in a successive water entry scenario of two spheres. Regarding the impact Froude number, three typical modes were introduced: (1) Steady mode: when Fr I > 33.0, the rear sphere falls into the opening cavity of the front sphere and then enters the water, (2) Transition mode: for 25.2 < Fr I 33.0, the rear sphere first contacts the splash formed by the front sphere and then enters the water, and (3) Perturbation mode: for 3.4 < Fr I 25.2, the second sphere enters the water after hitting the Worthington jet of the primary sphere. In 2023, Wang et al.88 conducted a study using a 3D numerical volume-of-fluid (VOF) approach and a six-degree of freedom solver to simulate the water entry of two spheres positioned side-by-side at varying lateral distances and time intervals. They validated their approach with the experimental results of Wang and Lyu.7 The findings of their study indicated that the time interval between the entries of the spheres does have an impact on the cavity dynamics, influencing the expansion and contraction processes. However, they noted that the influence of the time interval diminishes significantly when the lateral distance between the two spheres is increased.

Based on the literature review, it is evident that the manner in which projectiles enter the water (whether individually, in parallel, tandem, etc.) can have a strong effect on descent trajectories, cavity dynamics, fluid-structural integrity, and hydrodynamic forces. For instance, when a single sphere enters the water, the pinch-off time remains relatively constant across a wide range of release heights. However, in the case of parallel entry, the lateral distance between the objects and their surface wettability can have a significant impact on the pinch-off time. Therefore, the objective of this study is to experimentally investigate parallel water entry for spheres with identical size and material properties. It aims to examine the influence of different coating layers (hydrophobic and hydrophilic) and the lateral center-to-center distance between the spheres. The study focuses on analyzing the fundamental patterns of cavity formation, evolution, pinching, and the kinetics of projectiles. To effectively capture and analyze the dynamics of the water entry process, this study employs a high-speed photography system and an image processing technique. The paper is structured as follows: Sec. II provides a description of the experimental apparatuses and test procedures. Section III presents the qualitative outcomes, including the stages of cavity formation, evolution, and collapse as well as splash crown and Worthington jet behavior. Section IV presents the main quantitative results, which encompass the dimensions of the cavities and the dynamics of the projectiles. Finally, Sec. V concludes the paper by summarizing the main findings.

Figure 1 displays a schematic depiction of the water entry test rig and the associated coordinate system. The projectiles utilized in the experiment are steel spheres with a diameter of D s = 20 mm, a density of ρ s 7783 kg m 3, and a mass of m s 0.0326 kg. The original spheres have a smooth surface with a static contact angle of approximately 85°, indicating a hydrophilic behavior. In order to create hydrophobic spheres, some of the spheres are coated with a layer of ULTRA EVER DRY® TOP Coat produced by UltraTech International, Inc. This coating significantly increases the contact angle to around 165 170°, resulting in a superhydrophobic surface. In this paper, the abbreviations SHP and HPI refer to the spheres coated with the hydrophobic layer (superhydrophobic) and the uncoated spheres (hydrophilic), respectively. For testing purposes, an acrylic water tank with a cross-sectional area of 50 × 50 cm 2 and a depth of 150 cm is utilized. The tank is filled with water, with a water depth of approximately 130 cm. The experimental setup utilizes demineralized water, which possesses the following properties: a temperature range of T w 23 25 ° C, a density of ρ w 998 kg m 3, a dynamic viscosity of μ w 8.91 × 10 4 Pa s, and a surface tension of σ w 72.0 mN m 1.

FIG. 1.

(a) Schematic drawing of the test rig. (b) The measurement coordinate system.

FIG. 1.

(a) Schematic drawing of the test rig. (b) The measurement coordinate system.

Close modal

The release mechanism consists of a machined metallic plate with a row of small holes, an electromagnet, and a pushing spring. The plate is hinged to a fixed rod, and when it is attached to the electromagnet, it remains in a completely horizontal position. The dimension of the plate and the location of the holes have been designed so that the spheres impact the water surface in the center of the tank. The small holes are meticulously drilled to ensure specific center-to-center distances of the spheres ( l d), ranging from 1.0 D s to 5.0 D s in intervals of 0.5 D s. Upon deactivation of the electromagnet, the compressed spring forcefully propels the plate downward. This downward movement is fast enough to ensure that the spheres resting on the holes are simultaneously released. To ensure simultaneous impact and to validate the absence of sphere rotation caused by the release mechanism, the impact events of the two spheres were carefully examined using high-speed imaging. The entire release mechanism is mounted on an adjustable horizontal beam. This beam can be vertically moved using a motorized system, allowing for precise adjustment of the desired release height ( H R) above the water surface. When the spheres are released from the plate, they impact the water surface at a speed of U I = 2 g H R, where the acceleration due to gravity is represented by g = 9.81 m s 2. Therefore, for the range of release heights of H R = 15 95 cm used in these experiments, the impact velocities can be approximated as 1.72 U I 4.32 m s 1. Each simultaneous release test is conducted three times, and the data obtained from each test are averaged for analysis. This approach ensures reproducibility and accuracy in the measurements. Table I presents a summary of the key dimensional and dimensionless parameters required for the analysis.

TABLE I.

List of dimensional and non-dimensional variables of the conducted experiments.

Parameter Symbol Definition Range/values Unit
Sphere diameter  D s  ⋯  20  mm 
Release height  H R  ⋯  15 95  cm 
Sphere mass  m s  ⋯  32.6  g 
Center-to-center lateral distance  l d  ⋯  1.0 D s 5.0 D s  mm 
Impact velocity  U I  2 g H R  1.72 4.32  m s 1 
Impact Froude number  Fr I  U I / g D s  3.87 9.75  ⋯ 
Impact Weber number  We I  ρ w U I 2 D s / σ w  816 5167  ⋯ 
Bond number  Bo I  ρ w g D s 2 / σ w  54.4  ⋯ 
Impact Reynolds number  Re I  ρ w U I D s / μ w  38 477 96 833  ⋯ 
Impact Capillary number  Ca I  μ w U I / σ w  0.021 0.053  ⋯ 
Capillary length  λ c  σ w / ( ρ w g )  2.71  mm 
Solid–liquid density ratio  ϱ  ρ s / ρ w  7.79  ⋯ 
Parameter Symbol Definition Range/values Unit
Sphere diameter  D s  ⋯  20  mm 
Release height  H R  ⋯  15 95  cm 
Sphere mass  m s  ⋯  32.6  g 
Center-to-center lateral distance  l d  ⋯  1.0 D s 5.0 D s  mm 
Impact velocity  U I  2 g H R  1.72 4.32  m s 1 
Impact Froude number  Fr I  U I / g D s  3.87 9.75  ⋯ 
Impact Weber number  We I  ρ w U I 2 D s / σ w  816 5167  ⋯ 
Bond number  Bo I  ρ w g D s 2 / σ w  54.4  ⋯ 
Impact Reynolds number  Re I  ρ w U I D s / μ w  38 477 96 833  ⋯ 
Impact Capillary number  Ca I  μ w U I / σ w  0.021 0.053  ⋯ 
Capillary length  λ c  σ w / ( ρ w g )  2.71  mm 
Solid–liquid density ratio  ϱ  ρ s / ρ w  7.79  ⋯ 

To record the water entry processes, two high-speed cameras are employed. The first camera, a Photron FASTCam-SA3, operates at a frame rate of 2000 fps and possesses an image resolution of 1024 × 1024. It is specifically utilized for close-up capturing of the impact, splash formation, and cavity development. The second camera, a Kron Technologies CHRONOS-2.1HD, operates at a frame rate of 1512 fps and has an image resolution of 1280 × 1024. Its purpose is to capture wide-angle images from a distance to record the spheres' trajectories up to a water depth of approximately 50 cm. To achieve the desired shots, the following lenses are used: “Nikkor-35 mm-f/1.1,” “MicroNikkor-55 mm-f/2.8” for the close-up images, and “Nikkor-24–85 mm-f/2.8–4” for the wide-angle images. An LED light panel with sufficient contrast is mounted on the backside of the water tank. This ensures that the lighting conditions are optimal, resulting in high-quality images with clear contrast and detail resolution.

An in-house MATLAB script is employed for image processing to analyze various parameters in this study. The script is specifically developed to evaluate parameters such as the trajectory of the projectiles ( z s t in m), pinching time ( t p in ms), pinching depth ( L p in m), and cavity sizes. The script automatically analyzes the individual images of a water entry event and increases the contrast of the images to clearly identify the spheres and cavities. For impact scenarios where a clear distinction of the two spheres and the cavities is possible, the images are subsequently subdivided into distinct image regions to evaluate areas, centroid positions, and a bounding box for each object (connected component, i.e., sphere and attached cavity). The images are converted into black and white images, and continuous regions are identified to obtain the positions of the spheres and the cavities. In cases where the separation ratios of the spheres are low, individualized, not fully automated techniques are used to evaluate the sphere's positions.

In a subsequent step, a quintic smoothing spline is fitted to the captured trajectory data, using the algorithm described in Ref. 89 and used in Epps et al.,90 Techet and Truscott,91 and Truscott et al.92 This spline is used to determine sphere velocity [ V s t = z ̇ s in m s 1] and acceleration [ a s t = z ¨ s in m s 2] based on the derivatives of the spline. Here, t denotes the time, with t = 0 signifying the moment at which the projectile contacts the water surface. Additionally, the coefficient of total hydrodynamic force ( C F , s) is computed using the following equation:93 
(1)

In the realm of water entry, a range of events and phenomena can take place both above and below the water surface. These include the formation, evolution, and closure (pinch-off) of air entrainment cavities, the creation and sealing of splash crowns/domes, the generation of Worthington water jets, the presence of cavity wall waves/ripples, and the shedding of cavities, among others. The objective of this section is to provide a comprehensive analysis of qualitative and visual data related to the simultaneous impact of two spheres on the water surface. It explores specifically in detail the impact of the lateral center-to-center distance between the spheres and the surface wettability on the observed water entry phenomena. It is important to note that in this study, the term “symmetric” refers specifically to axisymmetry around the vertical axis of the spheres, as the photography and captured phenomena are two-dimensional. This should not be confused with symmetry with respect to the midplane between the spheres, which will be referred to as “mirrored.”

In Fig. 2 (Multimedia views), the chronophotography captures the water entry process of two hydrophobic spheres dropped from a height of 65 cm. The corresponding initial values of impact Weber, Reynolds, and Froude numbers are calculated as 3535, 80 089, and 8.1, respectively. The images are displayed with a time interval of Δ t = 15 ms, starting at t = 6 ms after the moment of impact and continuing up to 81 ms. The figure consists of three rows, each representing a different scenario. The first row illustrates the case where l d / D s is 5.0 (the farthest distance). The middle row displays the scenario with l d / D s = 2.0, while the last row showcases the case with l d / D s = 1.0 (the closest condition). From this figure, the following observations can be highlighted:
FIG. 2.

Chronophotography capturing the parallel water entry of two hydrophobic spheres dropped from a height of 65 cm ( We I = 3535 , Re I = 80 089 , Fr I = 8.1) at intervals of Δ t = 15 ms. (First row): the lateral center-to-center distance is 5.0 D s. (Second row): the lateral center-to-center distance is 2.0 D s. (Third row): the lateral center-to-center distance is 1.0 D s. Multimedia available online.

FIG. 2.

Chronophotography capturing the parallel water entry of two hydrophobic spheres dropped from a height of 65 cm ( We I = 3535 , Re I = 80 089 , Fr I = 8.1) at intervals of Δ t = 15 ms. (First row): the lateral center-to-center distance is 5.0 D s. (Second row): the lateral center-to-center distance is 2.0 D s. (Third row): the lateral center-to-center distance is 1.0 D s. Multimedia available online.

Close modal
It is evident that when the lateral center-to-center distance between two spheres entering the water is sufficiently large, the influence of the neighboring spheres on the cavity dynamics becomes negligible. As a result, the characteristics of the formed cavity closely resemble those observed in a single water entry process. In the scenario of l d / D s = 5.0, which represents the longest distance examined in the current study, this observation is particularly apparent, particularly in the initial stage of the impact. In this case, the formation of a nearly symmetric cavity is observed, and the pinch-off of the cavity occurs at a single point. To show the degree of symmetry in the cavities, a comparison is made between the mirrored left wall (red curve) and the original right wall (black curve) of the left air cavity in the figure corresponding to a time of 36 ms. Based on the Weber and Bond numbers, the cavity closure pattern can be classified as a surface-seal regime with a pinching event below the water surface. Pinch-off is observed at 66 ms. At that moment, the cavities have expanded and the distance between the facing walls of the cavities has decreased, leading to some asymmetry in their shapes, particularly for the upper cavities. Furthermore, as anticipated, the occurrence of pinching leads to the formation of ripples on the cavity walls and cavity shedding behind the spheres. These observations also provide evidence of the behavioral similarity, at least to some extent, between this configuration and the water entry of a single sphere. Rabbi83 also made a similar observation and remarked regarding the threshold value for the independent behavior of each sphere. He emphasized that for values of l d / D s < 4.0, the presence of a nearby sphere significantly influences the overall cavity formation. This suggests that l d / D s = 4.0 could potentially serve as a threshold value in determining the extent of the neighboring sphere's influence on cavity dynamics. Upon careful examination of the images recorded by Rabbi,83 it can be noted that even at l d / D s = 4.0, there is still a slight asymmetry in the cavities. However, 4.0 can be considered as an acceptable threshold value for the separation ratio. It is worth mentioning that the difference between his findings and the current observation at l d / D s = 5.0 could be attributed to the difference in the hydrophobic coatings used. Rabbi utilized Cytonix WX-2100 coating, which has a static contact angle of approximately 117°, significantly lower than the superhydrophobic coating employed in this study, which had a static contact angle of 165 170°. From a nondimensional perspective, Rabbi's experimental works83 were conducted within the range of approximately 380 We I 3100, 2.3 Fr I 6.5, and 25 Bo 200, which are similar to the range of parameters investigated in the current study (refer to Table I).
When the spheres are positioned in close proximity to each other, for instance with a separation ratio of l d / D s = 2.0, their interaction leads to the creation of two strongly asymmetrical cavities. The image demonstrates that the shape of the two cavities is remarkably similar, almost like a mirror image of each other. This observation suggests that the precise release mechanism used in the experiment minimizes any disturbances or irregularities in the cavity formation process. As the cavity undergoes necking, the sides facing each other display almost straight profiles, while the opposing sides exhibit curved surfaces. Such distorted cavity shapes indicate strong flow interaction at the boundary of the cavity surfaces. An analogous phenomenon, which justifies the shape of these cavities, can be achieved by replacing the neighboring sphere with an artificial wall with a slip boundary condition, positioned in the middle of the spheres. It should be noted that Rabbi83 conducted experiments on the water entry of a single sphere near a solid wall with a no-slip boundary condition. The study revealed similarities, such as cavity asymmetry, but also differences, particularly concerning the Worthington jet, when compared to the side-by-side water entry configuration. One intriguing observation for small lateral distances relates to the pinching event. As the sphere descends in a typical single water entry scenario, hydrostatic pressure and the radial flow field cause the cavity walls to collapse inward until they touch at a single point, i.e., the pinch-off point. Subsequently, the cavity is halved, forming an upper bowl-shaped free surface cavity and a lower air cavity attached to the sphere. However, in the current case of closely positioned spheres, the pinching event does not lead to the formation of two completely separated cavities. Instead, the two halves remain connected by a curved cavity shedding string or air tube. This phenomenon occurs because the presence of a uniform radial velocity field surrounding the air-entrainment cavity is essential for a clean pinch-off. However, in the case of asymmetric cavities, the velocity field is nonuniform, since the radial velocity in the region between the spheres is low, or nearly zero at small separation ratios. As a result, the cavity walls begin to wraparound each other, forming an air tube or string-like structure instead of separate, fully closed cavities [a schematic picture shows this interpretation in Fig. 3(a)]. It is worthwhile mentioning that the generation of the air tube apparently not only depends on the separation ratio but also the impact velocity. This claim originates from the presence of warped air tubes, which were also observed in the study conducted by Rabbi.83 In that study, the range of the Froude number for the tests (regarding the formulation in Table I) was 2.29 Fr I 6.55, which is similar to the range of the current study. However, this phenomenon was not observed or reported in the works of Wang and Lyu,7 whose experiments involved higher impact velocities, resulting in Froude numbers ranging from 37 to 40.
FIG. 3.

Schematic illustration demonstrating observations at low lateral sphere distances; (a) formation of cavity walls wrapping around and creating an air-tube. (b) Aerodynamic analogy to understand the phenomenon of spheres moving away from each other.

FIG. 3.

Schematic illustration demonstrating observations at low lateral sphere distances; (a) formation of cavity walls wrapping around and creating an air-tube. (b) Aerodynamic analogy to understand the phenomenon of spheres moving away from each other.

Close modal

For the smallest separation ratio ( l d / D s = 1.0), indicating that the two spheres are in direct contact with each other, significant distortions are observed in the shape of the cavities. After the spheres in contact with each other impact the water surface, a large merged air cavity is formed. This merged cavity is particularly evident in the initial time frames. However, as the spheres descend further into the water, they exert repelling forces onto each other, resulting in an ongoing increase in the separation distance between the spheres. Consequently, two distinct cavities form, each attached to its respective sphere. Although this repelling force and the subsequent separation also occur for other values of l d / D s, this is particularly pronounced in the case of l d / D s = 1.0. Figure 3(b) provides insight into the phenomenon of spheres moving apart employing an aerodynamic analogy. The figure illustrates that the shape of the sphere and the attached cavity resembles an asymmetric hydrofoil with one flat and one curved profile. The accelerated flow on the curved surface generates a low-pressure zone on the left side of the sphere (suction side) and the flow along the flat side creates a high-pressure zone on its right side (pressure side). This configuration creates a repelling force that acts from the pressure side toward the suction side, causing the spheres to move away from each other and increase their separation distance. An additional intriguing phenomenon, which is specifically observed in the case of l d / D s = 1.0, is the occurrence and the progression of the cavity-shedding string (air-tube) over time. To illustrate this phenomenon, Fig. 4 presents a selection of close-up images of the water entry event shown in the last row of Fig. 2. These images span from time t = 77 to 97 ms, with intervals of Δ t = 5 ms. Notably, around t = 87 ms, the air tubes come into contact with each other at a location close to the lower air cavities attached to the spheres. Subsequently, a third air tube forms between the existing strings. These three air tubes largely maintain their continuity until the time t = 97 ms. It should be pointed out that in the work of Rabbi,83 only a single merged air string for a specific state of l d / D s = 1.0 was observed and the presence of two or three separate air tubes was not mentioned. This discrepancy in the observations may be attributed to the higher interval of 15 ms between the presented images in Rabbi's study compared to the finer time resolution of 5 ms used in Fig. 4 of the present study. Another possible interpretation can refer to the difference in the experimental settings. The sphere's diameters and release heights used in this study lead to slender and long cavities attached to the spheres with a high aspect ratio, whereas the images provided in Rabbi's work show cavities with a lower aspect ratio, which are similar to a bluff body. Consequently, in the latter case, a more unsteady and turbulent wake is produced, which significantly distorts the strings and impedes the development of clearly identifiable separate strings. The more streamlined cavity shapes of this study experience a low degree of boundary layer separation and the strings emerging from the bubbles experience lower turbulent distortions, which allows for the formation of a third air tube further upward. It is worth noting that further comprehensive studies, including numerical simulations, are necessary to delve deeper into this phenomenon and gain a thorough understanding of its intricate details.

FIG. 4.

Chronophotography capturing the parallel water entry of two hydrophobic spheres at a drop height of H R = 65 cm and a lateral center-to-center distance of l d / D s = 1.0. The sequence of images covers a time range from t = 77 to 97 ms after the impact moment, with intervals of Δ t = 5 ms. The images demonstrate the progressive generation of a third air-tube over time.

FIG. 4.

Chronophotography capturing the parallel water entry of two hydrophobic spheres at a drop height of H R = 65 cm and a lateral center-to-center distance of l d / D s = 1.0. The sequence of images covers a time range from t = 77 to 97 ms after the impact moment, with intervals of Δ t = 5 ms. The images demonstrate the progressive generation of a third air-tube over time.

Close modal
In continuing the discussion and investigation of the simultaneous water entry of two hydrophobic spheres, it is worthwhile to shift focus to the events occurring above the water surface. In order to address this objective, Fig. 5 (Multimedia views) provides a visual representation of the progressive formation of the splash crown during the parallel water entry of the two hydrophobic spheres. The spheres are dropped from a height of H R = 25 cm, and the corresponding initial values of the impact Weber, Reynolds, and Froude numbers are calculated as 1360, 49 670, and 5.0, respectively. The images are presented at regular intervals of Δ t = 15 ms, commencing from t = 10 ms after the moment of impact and spanning until the time of 100 ms. At this particular release height, the cavity closure pattern exhibits a deep-seal regime, and the splash crown does not form a dome-like shape. Consequently, this setup offers an ideal opportunity to capture the Worthington jet above the water surface without it crashing into and hitting the splash curtain. Figure 5 consists of three columns, each representing a different scenario. The first column illustrates the case where l d / D s is 5.0 (the highest separation ratio). The second column displays the scenario with l d / D s = 2.5, while the third column shows the case with l d / D s = 1.0. From the figure, the following observations can be noticed:
FIG. 5.

Splash and Worthington jet chronophotography capturing the parallel water entry of two hydrophobic spheres dropped from a height of 25 cm ( We I = 1360.8 , Re I = 49 670 , Fr I = 5.0) at intervals of Δ t = 15 ms. (First column): the lateral center-to-center distance is 5.0 D s. (Second column): the lateral center-to-center distance is 2.5 D s. (Third column): the lateral center-to-center distance is 1.0 D s. Multimedia available online.

FIG. 5.

Splash and Worthington jet chronophotography capturing the parallel water entry of two hydrophobic spheres dropped from a height of 25 cm ( We I = 1360.8 , Re I = 49 670 , Fr I = 5.0) at intervals of Δ t = 15 ms. (First column): the lateral center-to-center distance is 5.0 D s. (Second column): the lateral center-to-center distance is 2.5 D s. (Third column): the lateral center-to-center distance is 1.0 D s. Multimedia available online.

Close modal
The observations from the current study suggest that the phenomena occurring above the water surface, specifically the events related to the splash crown formation, exhibit stronger deviations from symmetry compared to the phenomena observed underwater. In the initial stages of water entry, when the distance between the spheres is sufficiently large (represented in Fig. 5 by the first image 10 ms after impact for l d / D s = 5.0), the splash crowns exhibit a symmetric pattern and resemble the behavior observed in single water entry cases. However, over time, this symmetry becomes increasingly distorted. The surfaces facing each other start to move apart, resulting in a deviation from the initial symmetrical shapes. This interaction ultimately results in the formation of asymmetric splash curtains, as observed during the time interval between 25 and 70 ms. This distortion could be caused by slight asymmetries in the water movement after impact. Asymmetries of the airflow entering the cavities might also contribute to this effect. During the mentioned time interval, the splash curtains have risen significantly above the water surface. Due to the extremely thin nature of these raised splash sheets, even slight variations in the velocity field below the water surface or of the oncoming air flowing into the cavity may be able to cause significant distortions. The curvature of the splash sheets facing the midplane is influenced in such a way that, for example, at t 40 ms, the cavities' surfaces facing each other tilt inward. This might be caused by either air or water vortices forming close to the water surface. This conceptual understanding of the events is based on assumptions that require validation through numerical simulations. In the subsequent time interval between 70 and 85 ms, the splash crowns gradually collapse. In this same time span, pinching happens below the water surface, and consequently, Worthington jets form. Due to the absence of dome-shaped splashes, the generated Worthington jets have the ability to more effortlessly rise above the water surface. However, it is noteworthy that these jets exhibit a slight inclination and do not exit the water vertically, as presented in Fig. 5. This effect might be attributed to the asymmetry observed in the cavities before and at pinch-off (as illustrated by the comparison of cavity wall profiles in Fig. 6). The expansion of the cavity walls is slightly hindered toward the mid-central plane due to the proximity of the cavity walls that face each other. As a result, different curvatures occur at pinch-off, resulting in a slightly asymmetric pinching process. Consequently, the Worthington jet is not vertically oriented from its initiation, leading to the formation of an inclined Worthington jet above the water surface. It is noteworthy that the sequence of events described can also be observed in Rabbi's study.83 However, the current study achieved a better-mirrored symmetry about the midplane of the two splash crowns and the water jets formed. The figures presented in Rabbi's study83 show stronger distortions of the splash crown and a less pronounced, more distorted Worthington jet compared to the findings of the current study.
FIG. 6.

Comparison of the cavity wall profiles on the left (red curve) and right (blue curve) side for the right hydrophobic sphere in the SHP-SHP parallel water entry scenario ( H R = 25 cm, l d / D s = 5.0).

FIG. 6.

Comparison of the cavity wall profiles on the left (red curve) and right (blue curve) side for the right hydrophobic sphere in the SHP-SHP parallel water entry scenario ( H R = 25 cm, l d / D s = 5.0).

Close modal
Similar to the scenario with l d / D s = 5.0, the parallel water entry with l d / D s = 2.5 initially exhibits a symmetrical splash curtain at t = 10 ms. However, as the spheres penetrate deeper into the water, asymmetry becomes apparent in the splash structure. In contrast to the configuration with l d / D s = 5.0, the splash surfaces facing each other show a slight inward tilt. As the cavity expands, the inner splash sheets gradually approach each other. At t = 40 ms, they come into contact at a single location and merge. During the collapse of the individual splash crowns, the merged portion of the curtains forms a central planar jet, which at first remains at a roughly constant height. This can be observed in the images captured between t = 50 and 85 ms. At reduced separation ratios, a later pinch-off time is observed (which will be discussed further in the quantitative analysis in Sec. IV). In addition, as discussed previously, at low separation ratios, the upper and lower cavities remain connected by a thin air tube for a certain duration of time. This leads to a delayed Worthington jet formation. However, similar to the case with l d / D s = 5.0, the Worthington jets in this configuration are also oblique, albeit with a more pronounced inclination (see the images at t = 100 ms for both scenarios). The upward momentum of the Worthington jets may also lead to a rising central planar jet, as shown in Fig. 5; however, this detail has not been consistently observed in every repetition of this test case. The different jets collide with each other at a relatively short distance above the water surface.
When two hydrophobic spheres with a separation ratio of l d / D s = 1.0 fall and impact the water surface, the splash crowns that form during the early stages of the entry process touch each other and quickly merge, while the splash crown still exhibits a strong upward momentum. This results in the formation of a powerful central planar water jet rising far above the splash crown. This behavior is different from the observations made for l d / D s = 2.5, where the merging process takes place when the splash crowns have already begun to decline. At cases, l d / D s = 1.5 and 2.0 (not shown) merging also occurs before the splash crowns have reached their maximum height, and an elevated central planar jet can also be observed. In the case of l d / D s = 1.0, this planar jet concentrates into two distinct roughly cylindrical jets (jet-pair) at the borders of the splash crown. This behavior can be explained by the high surface tension at the edges of a planar jet and the subsequent retraction of these edges, a phenomenon described, e.g., by Song and Tryggvason94 and Sünderhauf et al.95 In scenarios with very low separation ratios (such as the pictures in the third row of Fig. 2 or in Fig. 4, although taken at a higher release height), the observations show the formation of a single merged upper cavity. As a result, only one distorted Worthington jet is generated, and it exhibits a nearly vertical trajectory. This Worthington jet ascends along the central region of the merged splash crown. Therefore, it hardly interacts with the cylindrical jet-pair at the borders of the splash crown, which can descend independently. Figure 7 provides a comprehensive schematic of the splash evolution, depicting the merged splash crown, planar jet formation, and the development of the jet-pair. Additionally, it highlights the interaction between the Worthington jet and the splash crown. Similar phenomena of splash curtain formation in the l d / D s = 1.0 scenario are also observed in Rabbi's investigation.83 The formation of a jet-pair out of the central planar jet can also be observed. Contrary to the experiments of this study, it seems like the central Worthington jet does interact with this jet-pair in Rabbi's experiments; however, the image progression presented in that work does not allow for a more detailed analysis.
FIG. 7.

A schematic conceptual drawing of the merged splash crown, splash planar jet, and jet-pair development (steps 1–3) as well as the interaction between the Worthington jet and splash crown (step 4), provided for the SHP-SHP parallel water entry scenario at low impact velocities (e.g., H R = 25 cm) and the separation ratio of l d / D s = 1.0.

FIG. 7.

A schematic conceptual drawing of the merged splash crown, splash planar jet, and jet-pair development (steps 1–3) as well as the interaction between the Worthington jet and splash crown (step 4), provided for the SHP-SHP parallel water entry scenario at low impact velocities (e.g., H R = 25 cm) and the separation ratio of l d / D s = 1.0.

Close modal

Here, to conclude the analysis of the splash phenomenon, Fig. 8 presents the events above the water surface for a case in the surface seal regime. This figure showcases the chronophotography of the splash and Worthington jet, capturing the simultaneous water entry of two hydrophobic spheres released from a height of 65 cm. This investigation encompasses three distinct scenarios, characterized by l d / D s ratios of 5.0 , 2.5, and 1.0. The illustrations show the emergence of a splash dome and the subsequent interaction of the Worthington jet with this dome. The initial splash dome for l d / D s = 5.0 and 2.5 is asymmetric. The subsequent surface sealing events produce splashes, which are not located in the center of the dome. After the surface sealing, a merged splash dome rises in the l d / D s = 2.5 scenario. The interaction of the Worthington jet with the splash dome produces complicated disruptions of the splash dome, which do not allow for a comprehensive analysis of the events and the Worthington jet angles in meticulous detail. The scenario l d / D s = 1.0 produces a central jet pair, as has been observed in the deep-seal regime discussed before. This jet pair persists during the surface sealing event and the subsequent rise of the merged splash dome, until it is finally completely distorted by a strong central Worthington jet (not shown in the figure).

FIG. 8.

Splash and Worthington jet chronophotography capturing the parallel water entry of two hydrophobic spheres dropped from a height of 65 cm ( We I = 3535 , Re I = 80 089 , Fr I = 8.1. (First column): the lateral center-to-center distance is 5.0 D s. (Second column): the lateral center-to-center distance is 2.5 D s. (Third column): the lateral center-to-center distance is 1.0 D s.

FIG. 8.

Splash and Worthington jet chronophotography capturing the parallel water entry of two hydrophobic spheres dropped from a height of 65 cm ( We I = 3535 , Re I = 80 089 , Fr I = 8.1. (First column): the lateral center-to-center distance is 5.0 D s. (Second column): the lateral center-to-center distance is 2.5 D s. (Third column): the lateral center-to-center distance is 1.0 D s.

Close modal
The parallel water entry of two identical spheres with different wettability surfaces unveils a series of intriguing phenomena. These phenomena become particularly fascinating when one sphere exhibits hydrophobic properties, while the other possesses hydrophilic properties. In such cases, the presence of mirror symmetry is absent, and each sphere exhibits a distinct kinetic behavior. It is indeed worth noting that, to the best of the current authors' knowledge, such a configuration has not been studied in the existing literature. As a result, the findings and reported results in this study can be considered novel and can contribute significantly to the understanding of the related phenomena. Figure 9 (Multimedia views) shows the cavity dynamics observed during the water entry process of one hydrophobic sphere (left) and one hydrophilic sphere (right) dropped from a height of 95 cm. The corresponding values of impact Weber, Reynolds, and Froude numbers are calculated as 5167, 96 824, and 9.7, respectively. The images are presented at regular intervals of Δ t = 15 ms, commencing from t = 8 ms subsequent to the moment of impact and continuing until the time reaches 83 ms. The figure is composed of three rows, each presenting a distinct scenario. The lateral distance ratios presented are identical to those of the SHP-SHP cases of Fig. 2, i.e., l d / D s is 5.0 for the first row, l d / D s = 2.0 for the second, and l d / D s = 1.0 for the third row. Based on the images presented in this figure, several noteworthy observations can be made as follows:
FIG. 9.

Chronophotography capturing the parallel water entry of one hydrophobic sphere and one hydrophilic sphere dropped from a height of 95 cm ( We I = 5167 , Re I = 96 824 , Fr I = 9.7) at intervals of Δ t = 15 ms. (First row): the lateral center-to-center distance is 5.0 D s. (Second row): the lateral center-to-center distance is 2.0 D s. (Third row): the lateral center-to-center distance is 1.0 D s. Multimedia available online. ; ;

FIG. 9.

Chronophotography capturing the parallel water entry of one hydrophobic sphere and one hydrophilic sphere dropped from a height of 95 cm ( We I = 5167 , Re I = 96 824 , Fr I = 9.7) at intervals of Δ t = 15 ms. (First row): the lateral center-to-center distance is 5.0 D s. (Second row): the lateral center-to-center distance is 2.0 D s. (Third row): the lateral center-to-center distance is 1.0 D s. Multimedia available online. ; ;

Close modal
Similar to the findings discussed in Sec. III A, when the lateral center-to-center distance is sufficiently large, such as in the case of l d / D s = 5.0, the behavior of the cavities and the kinetics of the spheres resemble those observed in the water entry of a single object. In this configuration, the hydrophobic sphere (left one) exhibits the formation of a nearly symmetrical cavity, a pinch-off event below the water surface, and a closure pattern of the cavity belonging to the surface-seal regime. As anticipated, the pinching event results in the formation of ripples on the cavity wall and the shedding of the cavity behind the spheres. Conversely, no cavity is formed behind the hydrophilic sphere (right one), and it descends in a nearly straight trajectory, which also agrees with the expected development. The observed behavior indicates that the hydrophilic sphere, which generates no cavity, travels at a lower velocity. This phenomenon has been investigated by Truscott et al.92 The velocity difference can be attributed to the formation and shedding of ring-like vortices in the wake of the hydrophilic sphere. In contrast, the presence of an air cavity behind the hydrophobic sphere suppresses the formation of significant vorticity. The difference in the sphere's deceleration can be explained by the kinetic energy contained in the vortex structures shed by the hydrophilic sphere. Due to the high-density ratio of 7.79, the unsteady vortex shedding only slightly impacts the path stability of the hydrophilic sphere, which becomes evident by its low deviation from the vertical path, which is indicated by a dashed vertical line. The unsteady vorticity has no significant impact on the cavity symmetry of the hydrophobic sphere and its trajectory, which is perfectly vertical.
The images in the middle row of Fig. 9 show the scenario where the spheres are positioned in close proximity to each other, specifically with a separation ratio of l d / D s = 2.0. The interaction between the two spheres results in the formation of an asymmetrical cavity behind the hydrophobic sphere (left sphere). Furthermore, the hydrophilic sphere exhibits a significant deviation from its vertical trajectory, veering toward the right. This deviation can be visually observed by referring to the vertical dashed line depicted in the figure. This movement is believed to be associated with asymmetrical vortex shedding, wherein the vortices shed by the object are drawn toward the nearby air cavity. The trajectory of the shed vorticity toward the cavity becomes evident through the deformation of the cavity, which is already noticeable at t = 38 ms and becomes increasingly pronounced as time progresses. This effect of the shed vorticity on the cavity shape behind the hydrophobic sphere is illustrated by small blue arrows in Fig. 9. This cavity experiences an upper pinch-off event, similar to the SHP-SHP case discussed before; however, a second (lower) pinch off-event due to the influence of the vortices shed from the hydrophilic sphere can be witnessed at t = 83 ms. Similar events cannot yet be witnessed in the experiments conducted for a slightly higher lateral spacing of l d / D s = 2.5 (not shown); however, for l d / D s = 1.5, a remarkably similar sequence of events (cavity deformation due to shedding vorticity and second pinch-off) takes place. One specific experiment, which has been influenced by an unwanted surface distortion of the hydrophilic sphere, can serve as a visualization of the events since a small amount of air bubbles remained attached to the top of the sphere after the impact. Due to these bubbles, the formation of a vortex, which is being attracted by the neighboring cavity, can be witnessed. A series of images of this event is shown in Fig. 10(a). The complexity of the flow phenomena involved in these cases needs further investigation, e.g., using particle image velocimetry or numerical flow simulations to interpret the interaction of the vortices and the cavity in more detail.
FIG. 10.

(a) Images of a specific experiment ( H R = 95 cm, l d / D s = 1.5), which has been influenced by an unwanted surface distortion of the hydrophilic sphere. Here, the vorticity shed by the hydrophilic sphere is visualized by small air bubbles trapped in the shed vortices. (b) The influence of the release height on the interaction between a hydrophilic sphere and the adjacent cavity. The release heights are 15, 35, 45, and 95 cm. The separation ratio between the spheres is 2.0. The images represent the pinching moment of the individual cases.

FIG. 10.

(a) Images of a specific experiment ( H R = 95 cm, l d / D s = 1.5), which has been influenced by an unwanted surface distortion of the hydrophilic sphere. Here, the vorticity shed by the hydrophilic sphere is visualized by small air bubbles trapped in the shed vortices. (b) The influence of the release height on the interaction between a hydrophilic sphere and the adjacent cavity. The release heights are 15, 35, 45, and 95 cm. The separation ratio between the spheres is 2.0. The images represent the pinching moment of the individual cases.

Close modal

The chronophotography presented in the last row of Fig. 9 illustrates the phenomena observed during the water entry of two in-contact spheres. The right sphere is hydrophilic, and according to expectations, the formation of an air entrainment cavity is deemed impossible at the given impact velocity. The images in the first and second rows of Fig. 9 support this notion, as no cavity is observed behind the hydrophilic sphere. However, in the third row, it becomes apparent that when the spheres are in close contact; i.e., at a separation ratio l d / D s = 1.0, the air from the main cavity formed by the hydrophobic left sphere partly encompasses the right hydrophilic sphere and, for the release height of the presented images, a smaller cavity remains attached at the hydrophilic right sphere. This can be observed in the images at t = 23 and 38 ms. After a few milliseconds ( t = 53 ms), this newly formed cavity separates from the main cavity and travels downward along with its respective sphere. This separation process is accompanied again by the formation of a narrow air tube, as can be observed in the image at t = 68 ms. At this stage, the cavity behind the left sphere takes on a shape resembling a long circular pipe, with a diameter determined by the leading sphere. This air pipe then pinches off, which results in the formation of a second narrow air tube, as can be observed at t = 83 ms. Interestingly, in this scenario, the previous observation of the hydrophilic sphere moving slower than the hydrophobic sphere is no longer true. Instead, the hydrophilic sphere now moves slightly faster. It is presumed that this behavior change can be attributed to four factors. First, the tear-shaped air cavity behind the right sphere suppresses the formation of ring-like vortices, resulting in a reduced drag on the combined sphere-cavity system. Second, differences in the separation location of the water flow on the sphere surface can lead to different pressure distributions on the sphere surface. Third, the hydrophilic sphere experiences a lower upward buoyancy force compared to the hydrophobic sphere due to its smaller cavity volume. Finally, the hydrophilic sphere travels a longer distance with a detached cavity from the water surface, minimizing the influence of surface tension forces between the meniscus and the sphere. In contrast, the hydrophobic sphere maintains an attached cavity that remains connected to the water surface for a significant duration, leading to a stronger impact of the mentioned surface tension force on its movement. It is important to note that the latter two factors will have a relatively small impact compared to the first two factors. It is worth noting that the occurrence of this double cavity regime depends on the impact velocity and is observed in scenarios where H R 55 cm. Conversely, in cases where H R 45 cm, a single cavity regime can be observed, similar to the scenarios with larger separation ratios described above, however without the occurrence of a second pinch-off event but a rather highly disturbed cavity. Upon analyzing the frame-by-frame images, it becomes evident that immediately after the impact, the cavity generated by the left hydrophobic sphere partially covers the adjacent hydrophilic sphere. Subsequently, the two spheres move slightly apart, resulting in water flow entering the gap between them. In cases where the sphere velocity is higher (i.e., H R 55 cm), the water flow entering the gap fails to completely cut off the air attached to the sphere, despite its hydrophilic surface. This implies that the water flow within the gap follows a more direct upward path, cutting through the entrained air at a location further away from the sphere's surface, leading to a cavity pinch-off event and an air cavity that remains attached to the hydrophilic sphere.

The regime presented and discussed in the second row of images in Fig. 9, where a strong influence of the hydrophilic sphere on the air-entrainment cavity of the hydrophobic sphere is observed, which results in the formation of a second pinching point, is explored further in the following. The effect of impact velocity is investigated and presented in Fig. 10(b). This figure highlights the interaction between the hydrophilic sphere and the cavity for four different release heights: H R =  15, 35, 45, and 95 cm. In all the images, the separation ratio of the spheres is l d / D s = 2.0. It is evident that, as the impact velocity increases, the strength of the vortices shed by the right sphere and their influence on the opposing cavity become more pronounced. This intensifies the deformation of the cavity shape and increases the likelihood of the formation of a second pinching point. A relationship between the minimum release height and the separation ratio regarding the occurrence of a second pinch-off event is shown in Table II. The data demonstrate that a second pinch-off is more likely to occur at lower separation ratios. When the hydrophilic sphere is located closer to the hydrophobic one, even the shedding of weaker vorticity can have a strong enough impact on the neighboring cavity to produce a second pinching event.

TABLE II.

Relationship between the separation ratio and minimum release height (including corresponding dimensionless numbers) for which a second pinch-off event occurs in SHP-HPI configurations.

Separation ratio, l d / D s Min. H R (cm) Min. Re I Min. We I
1.5a  35  58 770  1904 
2.0  55  73 672  2991 
2.5  95  96 824  5167 
Separation ratio, l d / D s Min. H R (cm) Min. Re I Min. We I
1.5a  35  58 770  1904 
2.0  55  73 672  2991 
2.5  95  96 824  5167 
a

Some repetitions at higher impact velocities did not produce a second pinch-off, a highly disturbed cavity was observed instead.

Section III A investigates the events beneath the water surface. As it is also valuable to examine the phenomena occurring above, this aspect is addressed in Fig. 11 (Multimedia views) and provides a visual representation of the splash crown formation during the simultaneous water entry of a hydrophobic and a hydrophilic sphere. Both spheres are released from a height of H R = 25 cm, resulting in initial values of the impact Weber, Reynolds, and Froude numbers being 1360, 49 670, and 5.0, respectively. The sequence of images is captured at regular intervals of Δ t = 15 ms, starting from t = 10 ms after the impact and continuing until 85 ms. This figure highlights two scenarios. The left column corresponds to the scenario where l d / D s = 2.5, and the right column corresponds to the scenario where l d / D s = 1.0. In the case of the separation ratio of 2.5, it becomes evident that both the hydrophobic and hydrophilic spheres exhibit behaviors similar to an isolated water entry scenario. Specifically, for the hydrophobic sphere, an approximately symmetric splash crown is observed. Conversely, the hydrophilic sphere on the right generates a single vertical water jet as the water film layer surrounding its surface is displaced, which aligns with its behavior in an independent water entry context. It should be noted that the hydrophobic sphere eventually creates an inclined Worthington jet, which indicates a slightly asymmetric cavity shape at pinch-off. In the case of a 1.0 separation ratio, similar to the events discussed for two hydrophobic spheres, the formation of the jet-pair is observed. However, this jet pair is not located in the center of the splash curtain and rises in an oblique direction. In addition to the formation of this jet-pair, another interesting phenomenon is observed. A transverse jet (instead of a vertical jet) emerges from the side of the splash crown facing the hydrophilic sphere. This jet is probably created due to the same movement of the water layer around the hydrophilic surface. The interaction with the splash crown of the left sphere does not permit a vertical jet development. Such a transverse jet can not only be observed in the deep seal regime presented in Fig. 11 but also in a surface seal regime. In the latter case, the jet emerges directly from the water surface and becomes weaker with increasing release height.

FIG. 11.

Splash and Worthington jet chronophotography capturing the parallel water entry of one hydrophobic and one hydrophilic sphere dropped from a height of 25 cm ( We I = 1360 , Re I = 49 670 , Fr I = 5.0) at intervals of Δ t = 15 ms. (First column): the lateral center-to-center distance is 2.5 D s. (Second column): the lateral center-to-center distance is 1.0 D s. Multimedia available online.

FIG. 11.

Splash and Worthington jet chronophotography capturing the parallel water entry of one hydrophobic and one hydrophilic sphere dropped from a height of 25 cm ( We I = 1360 , Re I = 49 670 , Fr I = 5.0) at intervals of Δ t = 15 ms. (First column): the lateral center-to-center distance is 2.5 D s. (Second column): the lateral center-to-center distance is 1.0 D s. Multimedia available online.

Close modal

In contrast to Sec. III, the primary focus of this section is to conduct a comprehensive analysis of quantitative data concerning the conducted experiments. The main objective is to examine in detail the influence of both the lateral center-to-center distance between the spheres at impact and the surface wettability pairing on various aspects, including the spheres' kinetics, the cavities' pinch-off time and depth, or the angles of Worthington jets. The data utilized in this analysis are obtained through the application of image processing techniques, allowing for a more precise and detailed investigation of the observed phenomena. The main steps of the employed algorithms are outlined in Sec. II. All experimental tests have been conducted with a minimum of three replications. The following results represent the average values of these replications. The square root of the variance, representing the deviation between the measured data and the average value, is used to quantify the uncertainty. The “error bar” symbol is employed to visually represent this uncertainty. However, to maintain clarity in the graphs, the uncertainty is explicitly emphasized in only one figure, namely, Fig. 16, while the remaining figures present the average values without error bars to avoid confusion.

The findings of the study demonstrate a substantial influence of the lateral center-to-center distance between the spheres and their interaction during parallel water entry on the values of pinch-off time ( t p) and pinch-off depth ( L p) of the cavities formed. Figures 12(a) and 12(b) showcase the variation of the dimensionless pinch-off depth ( L p / D s) and of the dimensionless pinch-off time ( t p U I / D s) in relation to the separation ratio ( l d / D s), respectively. An analysis of Fig. 12(a) shows that regardless of release height and surface wettability pairing, an increase in the separation ratio results in a reduction of the pinch-off depth. However, with increasing the separation ratio, the influence of the neighboring sphere on the pinch-off depth gradually diminishes, and the graphs tend to level off horizontally. These curves are expected to approach the same pinch-off depth observed in a single water entry scenario, particularly for very high separation ratios. It is worth noting that the inclination of the curves for the SHP-SHP scenario does not reach zero at l d / D s = 5.0, which concurs with the observation of a slight asymmetry in the cavities even at that separation ratio, as described in Sec. III. Furthermore, it is evident that for a given release height, the cavity closure of a hydrophobic sphere takes place at a shallower depth when it is in proximity to a neighboring hydrophilic sphere. This effect becomes less pronounced at higher release heights, so that for the largest release height considered in the current experiments (i.e., H R = 95 cm), the results for closure depth become nearly identical for both the SHP-SHP and SHP-HPI configurations.

FIG. 12.

(a) The dimensionless pinch-off depth ( L p / D s) as a function of the separation ratio ( l d / D s). (b) The dimensionless pinch-off time ( t p U I / D s) as a function of the separation ratio ( l d / D s). Horizontal green lines represent the semi-analytical relationship proposed by Duclaux et al.3 for a single water entry scenario.

FIG. 12.

(a) The dimensionless pinch-off depth ( L p / D s) as a function of the separation ratio ( l d / D s). (b) The dimensionless pinch-off time ( t p U I / D s) as a function of the separation ratio ( l d / D s). Horizontal green lines represent the semi-analytical relationship proposed by Duclaux et al.3 for a single water entry scenario.

Close modal

The graphs presented in Fig. 12(b) corroborate the findings of Fig. 12(a), indicating that an increase in the separation ratio leads to a decrease in the pinch-off time. This observation holds true regardless of the release height and surface wettability pairing. Moreover, it is evident that for high separation ratios, the pinch-off time stabilizes at a relatively constant value, aligning with the pinch-off time observed in a single-sphere water entry scenario. Once again, it is worth reiterating that the curves for the SHP-SHP scenario do not become fully horizontal since a slight influence of the neighboring sphere and the cavities remains even at l d / D s = 5.0. Figure 12(b) also includes three horizontal green lines that represent the semi-analytical dimensionless pinch-off time for single-sphere water entry, calculated using the formula proposed by Duclaux et al.,3 i.e., t p , = 2.06 D s / 2 g, for the cavity pinching time in an undisturbed deep-sealing regime (refer to the next paragraphs for an explanation of the notation chosen here). The comparison between these lines and the test results from this research provides valuable insight. First, it serves to validate the reliability of the release mechanism employed in the experiments, as the close alignment between the single sphere limit and the test results for large separation ratios suggests that the spheres were released without rotation or other distortions. Second, the comparison demonstrates that this relationship remains valid for single water entry of superhydrophobic spheres.

To further investigate the pinch-off phenomenon, Fig. 13(a) displays the dimensional pinch-off time (in seconds) for all cases presented in Fig. 12(b), plotted against the separation ratio. The findings indicate that the cavity closure time remains relatively consistent across different release heights, particularly for the SHP-SHP configuration. However, in the SHP-HPI cases, the unsteady vortex shedding process, which takes place behind the right hydrophilic sphere, can impact the left cavity. As a consequence, the data are not as consistent as in the SHP-SHP case, since differences in vortex sizes, strengths, and trajectories lead to larger scatter in the results. These observations suggest the possibility of developing a regression model to predict the pinch-off time in parallel water entry for the SHP-SHP configuration, similar to the theoretical relationship established for single water entry. The proposed regression model can be expressed as a simple modification of the relationship of Duclaux et al.3 and is given by the equation t p = a SHP SHP D s / 2 g. Here, a SHP SHP represents a coefficient that depends on the separation ratio l d / D s, whose value must converge toward 2.06 when the separation ratio approaches infinity. After performing a non-linear regression analysis using the SPSS software, the following formula for the coefficient a SHP SHP has been derived:
(2)
FIG. 13.

(a) The dimensional pinch-off time [ t p (s)] as a function of the separation ratio ( l d / D s) for various release heights and configurations SHP-SHP and SHP-HPI. (b) Comparison of the experimental results of Rabbi83 with the regression model proposed in Eq. (3). The red bars in the right figure mark the range of results obtained by Rabbi for different sphere's diameters and release heights. The solid green line represents the current regression model, while the dashed-dot green line corresponds to the potential flow model formula, i.e., Eq. (4) proposed by Rabbi83 for SHP-SHP configurations. The horizontal solid purple line represents the semi-analytical relationship proposed by Duclaux et al.3 for single water entry.

FIG. 13.

(a) The dimensional pinch-off time [ t p (s)] as a function of the separation ratio ( l d / D s) for various release heights and configurations SHP-SHP and SHP-HPI. (b) Comparison of the experimental results of Rabbi83 with the regression model proposed in Eq. (3). The red bars in the right figure mark the range of results obtained by Rabbi for different sphere's diameters and release heights. The solid green line represents the current regression model, while the dashed-dot green line corresponds to the potential flow model formula, i.e., Eq. (4) proposed by Rabbi83 for SHP-SHP configurations. The horizontal solid purple line represents the semi-analytical relationship proposed by Duclaux et al.3 for single water entry.

Close modal
The determination coefficient ( R 2) for this regression model is 0.981, indicating a strong fit to the data. The corresponding pinching-time ratio relative to the single water entry case, which is denoted by t p , , to represent an infinite lateral sphere distance, becomes
(3)
Regarding Eqs. (2) and (3), it is necessary to clarify that Rabbi83 introduced an equation in his study to predict the pinch-off time of parallel water entry using a potential flow model. His formula is expressed as follows:
(4)
Based on the regression relation and the formula proposed by Rabbi83 for the pinch-off time, a comparison was conducted, and the results are shown in Fig. 13(a) using a solid green line for the regression relation and a dashed-dot green line for the potential flow model formula. These graphs demonstrate an excellent fit of the data to the regression relation and highlight the deviation of the data from the potential flow model formula. Figure 13(b) presents a comparison of the relations given by Eqs. (3) and (4) with the experimental results extracted from Rabbi's work. These experiments were conducted using hydrophobic spheres of different diameters released from three different heights, and the red bars indicate the range of these experimental results. The empirical regression obtained from the experiments of this study provides a very good fit for Rabbi's experimental data. The graphs indicate that the potential flow model formula given by Eq. (4) tends to overestimate the pinch-off phenomenon across the entire range of lateral distance ratios. However, it captures a realistic trend within the range of 3.0 l d / D s 5.0. Notably, the model significantly overpredicts the lateral distance ratio at which the delay in pinch-off, compared to the single water entry scenario, becomes practically zero. The deviation between the potential flow model and both Rabbi's and the current experimental data can be attributed to the following factors: first, the assumption made in the derivation of Eq. (4) that the distance between the two spheres remains constant during the evolution of the cavities does not hold true in practical scenarios (see Sec. IV C for more details). Second, the potential flow model used in Eq. (4) cannot incorporate important factors such as drag, surface tension, weight, buoyancy, and other forces. These two issues seem to play a significant role in the cavity dynamics.

As mentioned in Sec. III A, the splash crown does not form a dome-like shape at low release heights. The absence of a dome-like shape provides the ideal opportunity to capture the Worthington jet since it does not crash into a splash curtain. Furthermore, in Fig. 5, it is observed that decreasing the lateral center-to-center distance in the SHP-SHP configuration results in an increase in the inclination angle of the Worthington jets ( θ). To quantify this observation, Fig. 14(a) illustrates the variation of θ as a function of time for two release heights ( H R = 15 and 25 cm) and three separation values of l d / D s = 3.0 , 4.0, and 5.0. The angle θ is measured manually as indicated in Fig. 14(a), and the provided values are calculated by averaging the measured values of the left and right Worthington jets. For separation values below a certain threshold ( l d / D s 2.0), the upper bowl-shaped cavities merge, resulting in the formation of a single chaotic Worthington jet instead of two separate jets (see Fig. 5). This merging makes it challenging to measure the inclination angle of the individual jets. One notable observation is the increasing angle of the jet over time, with a tendency to become more vertical. This behavior can be attributed to the reduction in the dimensions of the upper cavities and the reduction in the velocity of the water close to the surface as time progresses. As the upper cavities become smaller, their influence on each other decreases. This, together with the lower velocities in the water phase, leads to a more symmetrical shape. Consequently, the jets tend to align more vertically. Another observation is the correlation between the release height and the angles of the jets. It is observed that as the release height increases and the spheres impact the water surface at higher speeds, the jet trajectory becomes more oblique. This tilt can be attributed to the stronger asymmetry of the upper cavity resulting from the higher impact speeds. Finally, as expected, when the lateral center-to-center distance increases and the configuration tends more toward single water entry, the angle of the Worthington jets increases and the jets tend to be more vertical. This issue is also displayed using combined images in Fig. 14(b) for H R = 25 cm in eight selected time-frames. It is worth noting that Rabbi83 also reported a similar trend in the variation of jet angles with respect to l d / D s and release height. However, the development of the tilt angles over time is not addressed and the specific times at which the angles are evaluated are not specified in his study.

FIG. 14.

(a) Variation of Worthington jets ( θ) as a function of time for two release heights ( H R = 15 and 25 cm) and three separation ratios of l d / D s = 3.0 , 4.0, and 5.0. (b) The combined images compare the inclination of the jet for different separation ratios and H R = 25 cm.

FIG. 14.

(a) Variation of Worthington jets ( θ) as a function of time for two release heights ( H R = 15 and 25 cm) and three separation ratios of l d / D s = 3.0 , 4.0, and 5.0. (b) The combined images compare the inclination of the jet for different separation ratios and H R = 25 cm.

Close modal

As discussed in Sec. III A and supported by Fig. 3(b), it is observed that when two hydrophobic spheres impact the water at a low separation ratio, they exert horizontal forces on each other, resulting in an increase in the horizontal center-to-center distance between the spheres ( Δ x). To quantify this phenomenon, the dimensionless parameter Δ x / l d is plotted as a function of descent time (measured from the moment of impact, milliseconds) in Fig. 15. The release height is H R = 65 cm, and separation ratios are l d / D s = 1.0, 1.5, 2.0, 3.0, and 5.0. The figure demonstrates that for a separation ratio of l d / D s = 5.0, the spheres exhibit a slight tendency to move closer to each other over time. However, for lower separation ratios, an increase in the distance between the spheres is observed. This effect becomes more pronounced as the separation ratio decreases. In the case of the minimum separation ratio, the distance between the centers has more than doubled in size within the recorded trajectory range. The findings also indicate the presence of a threshold separation ratio of approximately 4.5. At values above this threshold, the spheres tend to attract each other, whereas, below it, they move farther apart.

FIG. 15.

Dimensionless horizontal center-to-center distance between the spheres as a function of descent time. The release height is H R = 65 cm, and separation ratios are l d / D s = 1.0, 1.5, 2.0, 3.0, and 5.0.

FIG. 15.

Dimensionless horizontal center-to-center distance between the spheres as a function of descent time. The release height is H R = 65 cm, and separation ratios are l d / D s = 1.0, 1.5, 2.0, 3.0, and 5.0.

Close modal

During the water entry process, the downward motion of the projectile is caused by the combined effects of gravity and kinetic inertia. Meanwhile, the motion is opposed by resistance forces, including the hydrodynamic drag force, buoyancy force, and forces due to surface tension. The dynamic behavior of the projectile during water entry is governed by the intricate interplay between these driving and resistance forces. However, as mentioned before, in the case of parallel water entry, specifically in the SHP-HPI configuration, additional factors come into play. The formation and shedding of ring-like vortices in the wake of the hydrophilic sphere have a significant impact, not only on the neighboring cavity but also on the kinetics of the spheres themselves. In Fig. 16, the kinetics of the SHP-HPI spheres released from a height of 95 cm (positioned at separation ratios l d / D s = 2.0, 2.5, 3.5, and 4.5) is presented as an illustrative example. The figure includes the spheres' vertical position over time z s t ( m ), the velocity V s t = z ̇ s ( m s 1), the acceleration a s t = z ¨ s( m s 2), and the total hydrodynamic force coefficient C F , s t. As mentioned in the first paragraph of Sec. IV, error bars are specifically illustrated for the case of l d / D s = 3.5. Highlighting the error bars provides a visual representation of the uncertainty associated with the measurements in that particular case representing the statistical contribution to the observed deviations. Systematic measurement errors, which are hard to quantify, e.g., due to camera positioning, lens distortion, etc., are not included and would lead to higher error bars. The results, especially those based on the second-time derivative of the vertical position data z s t, exhibit a certain waviness, which can be attributed to small errors in z s t. Based on the figure and the plotted graphs, several significant findings can be pointed out.

FIG. 16.

Kinetics of SHP-HPI spheres released from a height of H R = 95 cm (a) vertical position z s ( m ), (b) velocity V s ( m s 1 ), (c) acceleration a s ( m s 2 ), and (d) coefficient of total hydrodynamic force, C F , s.

FIG. 16.

Kinetics of SHP-HPI spheres released from a height of H R = 95 cm (a) vertical position z s ( m ), (b) velocity V s ( m s 1 ), (c) acceleration a s ( m s 2 ), and (d) coefficient of total hydrodynamic force, C F , s.

Close modal

Figures 16(a) and 16(b) indicate that in the range of these experiments, the spheres' separation ratio has a minimal effect on their vertical trajectory and descent velocity. However, it is noticeable that the hydrophilic sphere (HPI) experiences a slower descent and reaches its terminal velocity (around 2  . 0 m s 1) at a shallower depth compared to the hydrophobic sphere. Furthermore, details regarding these findings have been discussed in Sec. III B. The cases shown in Fig. 16 represent the cavity regimes shown in the first two rows of Fig. 9. The influence of the neighboring sphere on the cavity dynamics plays a minor role in the kinetics of steel spheres with a high solid-to-liquid density ratio. The low effect of the separation ratio indicates that the vertical velocity and vertical acceleration of the two spheres are similar to that observed in single water entry events.

Figure 16(c) showcases the vertical acceleration experienced by both spheres. Figure 16(d) presents the total hydrodynamic force coefficient calculated using Eq. (1). As anticipated, the deceleration of both spheres decreases over time, tending toward zero. However, it is notable that the hydrophilic sphere exhibits a higher deceleration in the first part of its descent compared to the hydrophobic counterpart. Consequently, the deceleration of the hydrophilic sphere reaches zero earlier, and it nearly reaches its terminal velocity within the measurement range of the experiments. The hydrophobic sphere still experiences significant deceleration at the end of the measurement range. The results for acceleration and the calculated force coefficient tend to be similar to those published by Truscott et al.92 for steel spheres with a solid-to-liquid density ratio of about 7.8. In the hydrophobic case, the inertia is so high that the dynamic effects of cavity pinching can hardly be detected in the deceleration curves. Similar to the observations reported by Truscott et al., the hydrophilic sphere exhibits higher initial deceleration and consequently higher resistive forces. The main reasons have been analyzed in detail by Truscott et al. Here, a brief summary regarding the main force components is provided, which differs slightly in the approach regarding the hydrophilic case, where the work of Mansoor et al.93 is considered. The deceleration rate is primarily influenced by the forces of added mass ( F a), buoyancy ( F b), unsteady vortex shedding ( F I), and the unsteady pressure fluctuations due to the expansion and collapse of cavities ( F c). The added mass can be estimated using the equation m a = C m ρ w V c, where V c represents the combined volume of the sphere and the attached cavity, and C m is the added mass coefficient. The value of C m depends on the flow pattern around the sphere and the attached cavity. For spheres moving without an attached cavity, C m is typically 0.5, while for a streamlined combination of sphere and cavity, a mean value of 0.672 is assumed over the entire descent. V c for a single descending sphere is generally lower than for a sphere traveling with an attached cavity. Consequently, it can be argued that F a , SHP > F a , HPI. The total buoyancy force also depends on V c (see Mansoor et al.93 for pre- and post-pinch-off expressions), which proves that F b , SHP > F b , HPI. For hydrophobic spheres, the attached cavity suppresses vortex shedding to a large extent, resulting in F I , SHP 0, however F c , SHP > 0 (especially at and after pinch-off). In contrast, for a non-cavity-forming hydrophilic sphere, the kinetic energy of the shed vorticity leads to F I , HPI 0, while the absence of a cavity results in F c , HPI 0. The main conclusion is consequently that the higher deceleration observed for the hydrophilic sphere can be attributed to the effect of F I , HPI, which describes the momentum transfer from the sphere motion to the unsteady vortex wake.

The experimental investigation conducted in this study focuses on the vertical parallel water entry of two spheres with identical geometric and material properties but differing surface wettability (hydrophilic or hydrophobic). In these tests, two different scenarios are emphasized: hydrophobic–hydrophobic (abbreviated as SHP-SHP) and hydrophobic–hydrophilic (abbreviated as SHP-HPI) side-by-side water entry. The spheres are simultaneously released from varying heights (ranging from 15 to 95 cm), resulting in a range of impact velocities of approximately 1.71 4.32 m s−1. The range for Bond, Froude, Weber, and Reynolds numbers listed in Table I indicates that the experimental tests were conducted within the deep and surface seal regimes. The lateral center-to-center distances between the spheres are also varied, ranging from 1.0 to 5.0 times the sphere's diameter. To analyze the water entry process, a high-speed photographing system is employed, and image processing techniques are utilized. The study examines various aspects, including the behavior of the air-entrainment cavity (shape, closure patterns, and air tubes), water flow features (Worthington jets and splashes), and sphere kinetics (position, velocity, acceleration, and forces). The parallel water entry of the two spheres in both wettability scenarios leads to asymmetry in the flow field surrounding the cavities, impacting various cavity parameters such as the pinching time, Worthington jet angle, the formation of air tubes between the upper and lower cavities, and the horizontal distance between the spheres during descent, among others. By decreasing the lateral center-to-center distances between the two spheres, these influences are further accentuated, revealing the intricate relationship between the spheres' positioning and the resulting water entry dynamics. This experimental study expands the understanding of water entry by investigating configurations that have not been tested thus far, in particular superhydrophobic coating in parallel water entry and pairing of superhydrophobic and hydrophilic spheres. Some specific and unique findings can be highlighted as follows:

  • SHP-SHP pairing:

    • When the separation ratio is 1.0, the air tubes come into contact with one another near the lower air cavities, which are attached to the spheres. Subsequently, a third air tube forms between the already existing structures.

    • When the separation ratio is 1.0, an interesting observation is made regarding the planar jet formed by the merging of the two splash crowns. It concentrates and splits into two distinct roughly cylindrical jets, referred to as “jet-pair,” located at the boundaries of the splash crown.

    • The study presents a regression formula for the pinch-off time of the parallel water entry of two hydrophobic spheres, which is compared to experiments by a previous study and a previously proposed theoretical formulation based on the potential flow model.

  • SHP-HPI pairing: Three main regimes regarding the cavity dynamics can be identified.

    • When separation ratios exceed a certain value, the influence of the neighboring spheres is negligible.

    • If the hydrophilic sphere comes closer to the hydrophobic sphere, its vortex wake disturbs the cavity behind the hydrophobic sphere, which can result in a second pinch-off event. The sphere's kinetics are hardly influenced by this disturbance and are akin to that of individual water entry events of hydrophobic or hydrophilic spheres, respectively.

    • At a separation ratio of 1.0 and high impact velocities, a smaller cavity detaches from the larger cavity behind the SHP-sphere and attaches to the HPI-sphere, moving along with it.

The authors would like to acknowledge Johannes Kepler University Linz for supporting this study. This project has received funding from the EU-Horizon 2020 Marie Skłodowska-Curie Individual Fellowship Program under the name of WE-EXPERTH (Grant No. 101022112), covering the period from 2022 to 2024. The authors would also like to express their appreciation to Professor Brenden Epps89 for publishing his Matlab code, “smoothspline.m version 5.0,” under a GNU general public license, which was utilized to determine the best smoothing spline fit for estimating the descent velocity and acceleration of the spheres.

The authors have no conflicts to disclose.

Pooria Akbarzadeh: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Resources (equal); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (equal). Michael Krieger: Conceptualization (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Supervision (equal); Writing – original draft (equal). Dominik Hofer: Formal analysis (equal); Investigation (equal); Software (equal); Visualization (equal). Maria Thumfart: Methodology (equal); Resources (equal). Philipp Gittler: Formal analysis (equal); Funding acquisition (equal); Project administration (equal); Supervision (equal).

The data that support the findings of this study are available from the corresponding author upon request. Movie files of this set of experiments will be made available on the Zenodo-platform (zenodo.org) and linked to the DOI of this article.

1.
J. W.
Glasheen
and
T. A.
McMahon
, “
Vertical water entry of disks at low Froude numbers
,”
Phys. Fluid
8
(
8
),
2078
2083
(
1996
).
2.
C.
Duez
,
C.
Ybert
,
C.
Clanet
, and
L.
Bocquet
, “
Making a splash with water repellency
,”
Nat. Phys.
3
(
3
),
180
183
(
2007
).
3.
V.
Duclaux
,
F.
Caillé
,
C.
Duez
,
C.
Ybert
,
L.
Bocquet
, and
C.
Clanet
, “
Dynamics of transient cavities
,”
J. Fluid Mech.
591
,
1
19
(
2007
).
4.
J. M.
Aristoff
,
T. T.
Truscott
,
A. H.
Techet
, and
J. W.
Bush
, “
The water entry of decelerating spheres
,”
Phys. Fluid
22
(
3
),
032102
(
2010
).
5.
S.
Gekle
and
J. M.
Gordillo
, “
Generation and breakup of Worthington jets after cavity collapse—Part 1: Jet formation
,”
J. Fluid Mech.
663
,
293
330
(
2010
).
6.
Y.
Kubota
and
O.
Mochizuki
, “
Influence of head shape of solid body plunging into water on splash formation
,”
J. Visualization
14
(
2
),
111
119
(
2011
).
7.
X.
Wang
and
X.
Lyu
, “
Experimental study on vertical water entry of twin spheres side-by-side
,”
Ocean Eng.
221
,
108508
(
2021
).
8.
K. G.
Bodily
,
S. J.
Carlson
, and
T. T.
Truscott
, “
The water entry of slender axisymmetric bodies
,”
Phys. Fluid
26
(
7
),
072108
(
2014
).
9.
J. O.
Marston
,
T. T.
Truscott
,
N. B.
Speirs
,
M. M.
Mansoor
, and
S. T.
Thoroddsen
, “
Crown sealing and buckling instability during water entry of spheres
,”
J. Fluid Mech.
794
,
506
529
(
2016
).
10.
M. R.
Moore
,
S. D.
Howison
,
J. R.
Ockendon
, and
J. M.
Oliver
, “
Three-dimensional oblique water-entry problems at small deadrise angles
,”
J. Fluid Mech.
711
,
259
280
(
2012
).
11.
S. J.
Kim
, “
Fluid dynamics in liquid entry and exit
,” Ph.D. thesis (
Virginia Polytechnic Institute and State University
,
2017
).
12.
T. T.
Truscott
,
B. P.
Epps
, and
J.
Belden
, “
Water entry of projectiles
,”
Annu. Rev. Fluid Mech.
46
,
355
378
(
2014
).
13.
P.
Akbarzadeh
,
M.
Norouzi
,
R.
Ghasemi
, and
S. Z.
Daghighi
, “
Experimental study on the entry of solid spheres into Newtonian and non-Newtonian fluids
,”
Phys. Fluid
34
(
3
),
033111
(
2022
).
14.
S. Y.
Sun
and
G. X.
Wu
, “
Local flow at plate edge during water entry
,”
Phys. Fluid
32
,
072103
(
2021
).
15.
O. M.
Faltinsen
,
Sea Loads on Ships and Offshore Structures
(
Cambridge University Press
,
Cambridge, UK
,
1990
).
16.
S.
Zhao
,
C.
Wei
, and
W.
Cong
, “
Numerical investigation of water entry of half hydrophilic and half hydrophobic spheres
,”
Math. Probl. Eng.
2016
,
5265818
.
17.
T. T.
Truscott
and
A. H.
Techet
, “
Water entry of spinning spheres
,”
J. Fluid Mech.
625
,
135
165
(
2009
).
18.
L.
Yang
,
Y.
Wei
,
C.
Wang
,
W.
Xia
,
J.
Li
,
Z.
Wang
, and
D.
Zhang
, “
Dynamics of the cavity evolution during vertical water entry of deformable spheres
,”
Phys. Fluids
33
,
065106
(
2021
).
19.
A.
May
, “
Water entry and the cavity-running behavior of missiles
,”
Report No. 75-2
(
Naval Surface Weapons Center, White Oak Laboratory
,
Maryland
,
1975
).
20.
O. M.
Faltinsen
and
R.
Zhao
, “
Water entry of ship sections and axisymmetric bodies
,” in AGARD FDP Workshop on High Speed Body Motion in Water (AGARD Report No. 827), Kiev, Ukraine (
AGARD
,
Neuilly sur Seine, France
,
1997
).
21.
C. H.
Chen
, “
Fatigue assessment for ship structure acted on nonlinear wave loads
,”
Key Eng. Mater.
353–358
,
2786
2789
(
2007
).
22.
Z.
Li
,
L.
Sun
,
X.
Yao
,
D.
Wang
, and
F.
Li
, “
Experimental study on cavity dynamics in high Froude number water entry for different nosed projectiles
,”
Appl. Ocean Res.
102
,
102305
(
2020
).
23.
Y.
Hirano
and
K.
Miura
, “
Water impact accelerations of axially symmetric bodies
,”
J. Spacecr. Rockets
7
,
762
764
(
1970
).
24.
Z.
Wu
,
C.
Zhang
,
J.
Wang
,
C.
Shen
,
L.
Yang
, and
L.
Ren
, “
Water entry of slender segmented projectile connected by spring
,”
Ocean Eng.
217
,
108016
(
2020
).
25.
J. M.
Aristoff
and
J. W. M.
Bush
, “
Water entry of small hydrophobic spheres
,”
J. Fluid Mech.
619
,
45
78
(
2009
).
26.
H. S.
Yoo
,
H. Y.
Choi
,
T. H.
Kim
, and
E. S.
Kim
, “
Effect of wettability on the water entry of spherical projectiles: Numerical analysis using smoothed particle hydrodynamics
,”
Phys. Fluid
12
,
035014
(
2022
).
27.
T.
Guillet
,
M.
Mouchet
,
J.
Belayachi
,
S.
Fay
,
D.
Colturi
,
P.
Lundstam
,
P.
Hosoi
,
C.
Clanet
, and
C.
Cohen
, “
The hydrodynamics of high diving
,”
Proceedings
49
(
1
),
73
(
2020
).
28.
J. V.
Ward
,
Aquatic Insect Ecology, Part 1: Biology, and Habitat
(
John Wiley
,
1992
).
29.
J. W.
Glasheen
and
T. A.
McMahon
, “
A hydrodynamical model of locomotion in the basilisk lizard
,”
Nature
380
,
340
342
(
1996
).
30.
T. M.
Wang
,
X. B.
Yang
,
J. H.
Liang
,
G. C.
Yao
, and
W. D.
Zhao
, “
CFD based investigation on the impact acceleration when a gannet impacts with water during plunge diving
,”
Bioinspiration Biomimetics
8
(
3
),
036006
(
2013
).
31.
S. I.
Sharker
,
S.
Holekamp
,
M. M.
Mansoor
,
F. E.
Fish
, and
T. T.
Truscott
, “
Water entry impact dynamics of diving birds
,”
Bioinspiration Biomimetics
14
(
5
),
056013
(
2019
).
32.
T.
Chen
,
W.
Huang
,
W.
Zhang
,
Y.
Qi
, and
Z.
Guo
, “
Experimental investigation on trajectory stability of high-speed water entry projectiles
,”
Ocean Eng.
175
,
16
24
(
2019
).
33.
D.
Gilbarg
and
R. A.
Anderson
, “
Influence of atmospheric pressure on the phenomena accompanying the entry of spheres into water
,”
J. Appl. Phys.
19
(
2
),
127
139
(
1948
).
34.
D. A.
Watson
,
C. J.
Souchik
,
M. P.
Weinberg
,
J. M.
Bom
, and
A. K.
Dickerson
, “
Making a splash with fabrics in hydrophilic sphere entry
,”
J. Fluid Struct.
94
,
102907
(
2020
).
35.
M.
Jamali
,
A.
Rostamijavanani
,
N. M.
Nouri
, and
M.
Navidbakhsh
, “
An experimental study of cavity and Worthington jet formations caused by a falling sphere into an oil film on water
,”
Appl. Ocean Res.
102
,
102319
(
2020
).
36.
C.
Shen
,
P.
Yu
,
T.
Wang
, and
N. Z.
Chen
, “
A CFD based Kriging model for predicting the impact force on the sphere bottom during the early-water entry
,”
Ocean Eng.
243
,
110304
(
2022
).
37.
J.
Wang
,
O. M.
Faltinsen
, and
C.
Lugni
, “
Unsteady hydrodynamic forces of solid objects vertically entering the water surface
,”
Phys. Fluid
31
(
2
),
027101
(
2019
).
38.
H.
Zeraatgar
,
J.
Malekmohammadi
,
M. J.
Javaherian
, and
H.
Moradi
, “
Sampling rate effect on wedge pressure record in water entry by experiment
,”
Ocean Eng.
179
,
51
58
(
2019
).
39.
A.
Shams
,
M.
Jalalisendi
, and
M.
Porfiri
, “
Experiments on the water entry of asymmetric wedges using particle image velocimetry
,”
Phys. Fluid
27
(
2
),
027103
(
2015
).
40.
G. D.
Xu
,
W. Y.
Duan
, and
G. X.
Wu
, “
Numerical simulation of water entry of a cone in free-fall motion
,”
Q. J. Mech. Appl. Math.
64
(
3
),
265
285
(
2011
).
41.
J. F.
Louf
,
B.
Chang
,
J.
Eshraghi
,
A.
Mituniewicz
,
P. P.
Vlachos
, and
S.
Jung
, “
Cavity ripple dynamics after pinch-off
,”
J. Fluid Mech.
850
,
611
623
(
2018
).
42.
T.
Sun
,
C.
Shi
,
G.
Zhang
,
B.
Zhou
, and
H.
Wang
, “
Cavity dynamics of vertical water entry of a truncated cone–cylinder body with different angles of attack
,”
Phys. Fluid
33
,
055129
(
2021
).
43.
Y. T.
Sui
,
S.
Li
, and
F. R.
Ming
, “
An experimental study of the water entry trajectories of truncated cone projectiles: The influence of nose parameters
,”
Phys. Fluid
34
,
052102
(
2022
).
44.
C.
Chen
,
X.
Yuan
,
X.
Liu
, and
J.
Dang
, “
Experimental and numerical study on the oblique water-entry impact of a cavitating vehicle with a disk cavitator
,”
Int. J. Nav. Archit.
11
(
1
),
482
494
(
2019
).
45.
Z. H.
Ma
,
D. M.
Causon
,
L.
Qian
,
C. G.
Mingham
,
T.
Mai
,
D.
Greaves
, and
A.
Raby
, “
Pure and aerated water entry of a flat plate
,”
Phys. Fluid
28
,
016104
(
2016
).
46.
A.
Iafrati
, “
Experimental investigation of the water entry of a rectangular plate at high horizontal velocity
,”
J. Fluid Mech.
799
,
637
672
(
2016
).
47.
S.
Tödter
,
O.
El-Moctar
,
J.
Neugebauer
, and
T. E.
Schellin
, “
Experimentally measured hydroelastic effects on impact-induced loads during flat water entry and related uncertainties
,”
J. Offshore Mech. Arct. Eng.
142
(
1
),
011604
(
2020
).
48.
D.
Liu
and
G.
Yao
, “
Vibration and stability analysis on the water entry process of a thin plate
,”
Ocean Eng.
244
,
110445
(
2022
).
49.
P.
Sun
,
A. M.
Zhang
,
S.
Marrone
, and
F.
Ming
, “
An accurate and efficient SPH modeling of the water entry of circular cylinders
,”
Appl. Ocean Res.
72
,
60
75
(
2018
).
50.
Y.
Chang
and
A. Y.
Tong
, “
A numerical study on water entry of cylindrical projectiles
,”
Phys. Fluid
33
,
093304
(
2021
).
51.
B.
Zhou
,
H.
Liu
,
Z.
Wu
,
X.
Han
,
T.
Sun
, and
G.
Zhang
, “
Numerical study of vertical water entry of cylinder under the influence of wind and current
,”
Phys. Fluid
33
(
3
),
033330
(
2021
).
52.
A. Z.
Chaudhry
,
Y.
Shi
, and
G.
Pan
, “
Study on the oblique water entry impact performance of AUV under different launch conditions based on coupled FEM-ALE method
,”
AIP Adv.
10
(
11
),
115020
(
2020
).
53.
Y.
Shi
,
Y.
Hua
, and
G.
Pan
, “
Experimental study on the trajectory of projectile water entry with asymmetric nose shape
,”
Phys. Fluid
32
,
122119
(
2020
).
54.
Y.
Hou
,
Z.
Huang
,
Z.
Chen
,
Z.
Guo
, and
L.
Han
, “
Different closure patterns of the hollow cylinder cavities with various water-entry velocities
,”
Ocean Eng.
221
,
108526
(
2021
).
55.
Y.
Hou
,
Z.
Huang
,
Z.
Chen
,
Z.
Guo
, and
Y.
Luo
, “
Investigations on the vertical water-entry of a hollow cylinder with deep-closure pattern
,”
Ocean Eng.
190
,
106426
(
2019
).
56.
B.
Zhou
,
H.
Liu
,
Y.
Wang
,
Z.
Wu
,
X.
Han
, and
W. M.
Gho
, “
Numerical investigation on the cavity dynamics and multiphase flow field evolution for water entry of vertical cylindrical shell
,”
J. Fluid Struct.
103
,
103268
(
2021
).
57.
M.
Jalalisendi
and
M.
Porfiri
, “
Water entry of cylindrical shells: Theory and experiments
,”
AIAA J.
56
(
11
),
4500
–4514 (
2018
).
58.
R.
Panciroli
,
S.
Ubertini
,
G.
Minak
, and
E.
Jannelli
, “
Experiments on the dynamics of flexible cylindrical shells impacting on a water surface
,”
Exp. Mech.
55
(
8
),
1537
1550
(
2015
).
59.
I. U.
Vakarelski
,
E.
Klaseboer
,
A.
Jetly
,
M. M.
Mansoor
,
A. A.
Aguirre-Pablo
,
D. Y. C.
Chan
, and
S. T.
Thoroddsen
, “
Self-determined shapes and velocities of giant near-zero drag gas cavities
,”
Sci. Adv.
3
(
9
),
e1701558
(
2017
).
60.
K. E.
Crandell
,
R. O.
Howe
, and
P. L.
Falkingham
, “
Repeated evolution of drag reduction at the air-water interface in diving
,”
J. R. Soc. Interface
16
(
154
),
20190125
(
2019
).
61.
A.
Mehri
and
P.
Akbarzadeh
, “
Hydrodynamic characteristics of heated/non-heated and grooved/un-grooved spheres during free-surface water entry
,”
J. Fluid Struct.
97
,
103100
(
2020
).
62.
A.
Mehri
and
P.
Akbarzadeh
, “
Water entry of grooved spheres: Effect of the number of grooves and impact velocity
,”
J. Fluid Struct.
100
,
103198
(
2021
).
63.
H.
Shokri
and
P.
Akbarzadeh
, “
Experimental investigation of water entry of dimpled spheres
,”
Ocean Eng.
250
,
110992
(
2022
).
64.
B.
Holfeld
,
F.
Maier
,
M.
Izzo
, and
S.
Dinardo
, “
Spatial high-speed-imaging of projectile impacts into fluids in microgravity
,”
Microgravity Sci. Technol.
21
,
73
77
(
2009
).
65.
S. D.
Guleria
,
A.
Dhar
, and
D. V.
Patil
, “
Experimental insights on the water entry of hydrophobic sphere
,”
Phys. Fluid
33
,
102109
(
2021
).
66.
M.
Greenhow
and
S.
Moyo
, “
Water entry and exit of horizontal circular cylinders
,”
Philos. Trans. R. Soc. A
355
,
551
(
1997
).
67.
T.
Ye
,
D.
Pan
,
C.
Huang
, and
M.
Liu
, “
Smoothed particle hydrodynamics (SPH) for complex fluid flows: Recent developments in methodology and applications
,”
Phys. Fluid
31
,
011301
(
2019
).
68.
M. M.
Zdravkovich
, “
Review of flow interference between two circular cylinders in various arrangements
,”
ASME. J. Fluids Eng.
99
(
4
),
618
633
(
1977
).
69.
M.
Mahbub Alam
,
M.
Moriya
, and
H.
Sakamoto
, “
Aerodynamic characteristics of two side-by-side circular cylinders and application of wavelet analysis on the switching phenomenon
,”
J. Fluid Struct.
18
(
3–4
),
325
346
(
2003
).
70.
L.
Chen
,
J. Y.
Tu
, and
G. H.
Yeoh
, “
Numerical simulation of turbulent wake flows behind two side-by-side cylinders
,”
J. Fluid Struct.
18
(
3–4
),
387
403
(
2003
).
71.
M.
Mahbub Alam
and
Y.
Zhou
, “
Flow around two side-by-side closely spaced circular cylinders
,”
J. Fluid Struct.
23
(
5
),
799
805
(
2007
).
72.
D.
Yoon
and
K.
Yang
, “
Flow-induced forces on two nearby spheres
,”
Phys. Fluids
19
(
9
),
98103
(
2007
).
73.
P.
Sooraj
,
M. H.
Khan
,
A.
Sharma
, and
A.
Agrawal
, “
Wake analysis and regimes for flow around three side-by-side cylinders
,”
Exp. Therm. Fluid Sci.
104
,
76
88
(
2019
).
74.
X. D.
Bai
,
W.
Zhang
, and
Y.
Wang
, “
Deflected oscillatory wake pattern behind two side-by-side circular cylinders
,”
Ocean Eng.
197
,
106847
(
2020
).
75.
W.
Wang
and
Y.
Wang
, “
Influence of parameters of twin cylinders on hydrodynamic during synchronous water entry
,”
Chin. J. Hydrodyn.
25
(
1
),
23
30
(
2010
).
76.
P.
Ern
and
N.
Brosse
, “
Interaction of two axisymmetric bodies falling side by side at moderate Reynolds numbers
,”
J. Fluid Mech.
741
,
R6
(
2014
).
77.
R.
Yousefnezhad
and
H.
Zeraatgar
, “
A parametric study on water-entry of a twin wedge by boundary element method
,”
J. Mar. Sci. Technol.
19
,
314
326
(
2014
).
78.
D.
Yu
, “
Water entry cavitation evolution and motion characteristics of double revolving body in tandem
,” M.S. thesis (
Harbin Institute of Technology
,
China
,
2018
).
79.
J.
Lu
,
Y.
Wei
,
C.
Wang
,
L.
Lu
, and
H.
Xu
, “
Experimental study on cavity evolution characteristics in the water-entry process of parallel cylinders
,”
Chin. J. Theor. Appl. Mech.
51
(
2
),
1
12
(
2019
).
80.
H.
Yun
,
X.
Lyu
, and
Z.
Wie
, “
Experimental study on oblique water entry of two tandem spheres with collision effect
,”
J. Visualization
23
,
49
59
(
2020
).
81.
H.
Yun
,
X.
Lyu
, and
Z.
Wie
, “
Experimental study on vertical water entry of two tandem spheres
,”
Ocean Eng.
201
,
107143
(
2020
).
82.
R.
Rabbi
,
N. B.
Speirs
,
A.
Kiyama
,
J.
Belden
, and
T. T.
Truscott
, “
Impact force reduction by consecutive water entry of spheres
,”
J. Fluid Mech.
915
,
A55
(
2021
).
83.
R.
Rabbi
, “
Impacts in fluids: Forces, scaling laws and bouncing modes in projectile impacts on water pools and droplets
,” Ph.D. thesis (
Utah State University
,
Logan, UT
,
2021
).
84.
X.
Lyu
,
H.
Yun
, and
Z.
Wie
, “
Influence of time interval on the water entry of two spheres in tandem configuration
,”
Exp. Fluids
62
,
222
(
2021
).
85.
L.
Lu
,
C.
Wang
,
Q.
Li
, and
P. K.
Sahoo
, “
Numerical investigation of water-entry characteristics of high-speed parallel projectiles
,”
Int. J. Nav. Archit. Ocean Eng.
13
,
450
465
(
2021
).
86.
L.
Lu
,
X.
Yan
,
Q.
Li
,
C.
Wang
, and
K.
Shen
, “
Numerical study on the water-entry of asynchronous parallel projectiles at a high vertical entry speed
,”
Ocean Eng.
250
,
111026
(
2022
).
87.
X.
Lyu
,
X.
Wang
,
H.
Yun
, and
Z.
Chen
, “
On water-entry modes of the latter sphere in tandem configuration with two spheres
,”
J. Fluid Struct.
112
,
103601
(
2022
).
88.
X.
Wang
,
X.
Lyu
,
R.
Sun
, and
D.
Tang
, “
Numerical study on the cavity dynamics for vertical water entries of twin spheres
,”
Def. Technol.
(published online 2023).
89.
B. P.
Epps
, “
An impulse framework for hydrodynamic force analysis: Fish propulsion, water entry of spheres, and marine propellers
,” Ph.D. thesis (
MIT
,
2010
).
90.
B. P.
Epps
,
T. T.
Truscott
, and
A. H.
Techet
, “
Evaluating derivatives of experimental data using smoothing splines
,” in
Proceedings of Mathematical Methods in Engineering International Symposium (MMEI), Lisbon, Portugal, 21–24 October
(
2010
).
91.
A. H.
Techet
and
T. T.
Truscott
, “
Water entry of spinning hydrophobic and hydrophilic spheres
,”
J. Fluid Struct.
27
(
5–6
),
716
726
(
2011
).
92.
T. T.
Truscott
,
B. P.
Epps
, and
A. H.
Techet
, “
Unsteady forces on spheres during free surface water entry
,”
J. Fluid Mech.
704
,
173
210
(
2012
).
93.
M. M.
Mansoor
,
I. U.
Vakarelski
,
J. O.
Marston
,
T.
Truscott
, and
S. T.
Thoroddsen
, “
Stable-streamlined and helical cavities following the impact of Leidenfrost spheres
,”
J. Fluid Mech.
823
,
716
754
(
2017
).
94.
M.
Song
and
G.
Tryggvason
, “
The formation of thick borders on an initially stationary fluid sheet
,”
Phys. Fluids
11
(
9
),
2487
2493
(
1999
).
95.
G.
Sünderhauf
,
H.
Raszillier
, and
G.
Tryggvason
, “
The retraction of the edge of a planar liquid sheet
,”
Phys. Fluids
14
(
1
),
198
(
2002
).