Next Article in Journal
Different Methods for Estimating Default Parameters of Alpha Power-Transformed Power Distributions Using Record-Breaking Data
Next Article in Special Issue
Research, Application and Future Prospect of Mode Decomposition in Fluid Mechanics
Previous Article in Journal
Identifying the Level of Symmetrization of Reaction Time According to Manual Lateralization between Team Sports Athletes, Individual Sports Athletes, and Non-Athletes
Previous Article in Special Issue
A Review of the Development and Research Status of Symmetrical Diaphragm Pumps
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Non-Axisymmetric Bouncing Dynamics on a Moving Superhydrophobic Surface

1
School of Energy and Power Engineering, University of Shanghai for Science and Technology, Shanghai 200093, China
2
School of Aerospace Engineering and Applied Mechanics, Tongji University, Shanghai 200092, China
3
School of Air Transportation, Shanghai University of Engineering Science, Shanghai 201620, China
4
School of Aeronautics and Astronautics, Shanghai Jiao Tong University, Shanghai 200030, China
5
Key Laboratory (Fluid Machinery and Engineering Research Base) of Sichuan Province, Xihua University, Chengdu 610039, China
*
Author to whom correspondence should be addressed.
Symmetry 2024, 16(1), 29; https://doi.org/10.3390/sym16010029
Submission received: 27 November 2023 / Revised: 12 December 2023 / Accepted: 20 December 2023 / Published: 25 December 2023
(This article belongs to the Special Issue Symmetry in Micro/Nanofluid and Fluid Flow)

Abstract

:
The phenomenon of droplet impact on moving surfaces is widely observed in fields such as transportation, rotating machinery, and inkjet printing. Droplets exhibit non-axisymmetric behavior due to the motion of solid surfaces which significantly determines core parameters such as contact time, maximum spreading radius, and bounding velocity, thereby affecting the efficiency of related applications. In this study, we focus on the kinetics and morphology of the non-axisymmetric bouncing behaviors for droplets impacting on a moving superhydrophobic surface (SHPS) within the normal (Wen) and tangential (Wet) Weber numbers. Considering the influences of the moving surface on the contact area and contact time, the previous scaling formula for the horizontal velocity of droplets has been improved. Based on the velocity superposition hypothesis, we establish a theoretical model for the ratio of the maximum spreading radius at both ends depending on Wen and Wet. This research provides both experimental and theoretical evidence for understanding and controlling the non-axisymmetric behavior of droplets impacting on moving surfaces.

1. Introduction

During the process of liquid droplet impact, various phenomena such as stacking, rebounding, and splashing can occur under the influence of the droplet properties and surface characteristics [1,2,3]. The rebound process refers to the lateral spreading of the droplet, with the conversion of kinetic energy to surface energy, gradually reaching the maximum spreading length Dmax, after which the droplet recoils towards the center until it reaches a certain degree of contraction. Subsequently, the droplet rebounds from the surface [4,5,6]. The phenomenon of droplet impact has been extensively studied both numerically and experimentally [7,8]. The rebound behavior of liquid droplets is influenced by various factors such as droplet properties [9,10], droplet radius and velocity [11,12], surface properties [13,14], and incident angle [15,16].
In previous studies on droplet impact, the predominant focus was on stationary surfaces, and the maximum spreading length of the impacting droplet was dependent on the normal Weber number Wen, especially in cases involving low viscosity and limited wetting ability [17]. When a droplet impacts on a stationary flat surface, the contact time was found to be contingent upon the radius and independent of the impact velocity [18]. In this context, contact time is defined as the duration from the first contact of the droplet with the surface to its complete detachment. Research into contact time has garnered practical applications in the field of self-cleaning surfaces [19]. Here, the design of macro-textures and microstructures can effectively reduce the contact time [20,21].
However, in practical applications, droplet impact phenomena mostly occur on moving surfaces, such as raindrops impacting aircraft wings and wind turbine blades [22], the size of the droplets influencing the process of extinguishing fires [23], and the directional transport of droplets in inkjet printing [24]. A few studies have found that on the SHPSs, as the speed of the moving surface increases, elevated surface velocities amplify the extension of impacting droplets, consequently leading to a reduction in contact time [25,26]. Similar findings have also been observed on moving surfaces with macroscopic structures and wetting patterns [27].
These conclusions mostly analyze droplets from a kinematic perspective, focusing on droplet spreading and contact time. There are still some conclusions related to momentum. Some studies suggest the main cause of momentum transfer is the aerodynamic Leidenfrost effect [28]. This phenomenon refers to the thin air film formed between the droplet and the moving surface, which is the primary factor generating viscous forces on the droplet. By analyzing the forces acting on the thin air film between droplet and surface, scaling relationships can be derived to summarize the contact time, spreading ratio, and horizontal distance. These dynamic conclusions have been well validated [29]. However, in these experiments, the presence of an air film between the liquid droplet and the surface was not distinctly observed. Droplets will inevitably come into contact with the surface, and the reason for this difference may be due to insufficient surface velocity to generate a noticeable air layer. We tend to the view that the transfer of horizontal momentum is caused by the viscous boundary layer developing in the liquid [30]. Throughout the impact process, energy dissipation due to viscous losses is observed. The fundamental physical principle of non-axisymmetric bouncing dynamics is the transfer of horizontal momentum during the impact process, gradually shifting from the bottom to the top of the droplet. Based on this, a formula related to the horizontal velocity of droplets has been proposed [31]; while the feasibility of this velocity formula has been well validated, there are still some sections for improvement. For example, in the derivation of this formula, the spreading region of the droplet is considered to be circular, which is accurate when impacting a stationary surface but imprecise on a moving surface, where the spreading region is elongated [29,32]. Furthermore, the contact time of the droplet is also assumed to be the same as the contact time on a stationary surface τs, whereas on a moving surface the contact time τm is shorter than τs [25,26].
Moreover, there is still a lack of satisfactory solutions regarding how the spreading range of droplets is influenced by surface motion. In this study, the fundamental dynamic behaviors of droplet impact on moving surfaces are focused on being revealed, and a more accurate correction has been made to the scaling formula for the droplet horizontal velocity, as well as an analysis of the asymmetric evolution of the spreading radius at both ends of the droplet over time caused by surface motion.

2. Methodology

We conducted experimental studies on SHPSs with horizontal movement speeds of Vt = 0–2.2 m/s (corresponding to Wet = 0–116.08) using water droplets with impact velocities of Vn = 0.48–1.70 m/s (corresponding to Wen = 5.69–71.26), where Wet = ρVt2D0/γ and Wen = ρVn2D0/γ are defined as the normal and tangential Weber number, respectively. Water density was ρ ≈ 1000 kg/m3, surface tension coefficient γ ≈ 0.073 N/m, and D0 = 1.8 ± 0.1 mm was the initial diameter of the water droplets. Figure 1a shows the schematic diagram of the experimental apparatus, where the copper plate was polished, ultrasonically cleaned, and coated with an organic reagent three times. After drying at room temperature for 30 min, a solution mainly composed of nano-silica particles and silicone resin was sprayed, and the static contact angle of water droplets on the surface was about 160° (Figure 1a). The surface was fixed to a rotating motor to obtain the horizontal speed, and the water droplets were pushed by the injection pump (LSP01-3A) and dropped onto the SHPS under the action of gravity after separating from the needle tip. The impact velocity of the droplets was adjusted by varying the distance between the needle tip and the surface. The position of the high-speed camera (GX-8E, NAC), the point of droplet impact, and the center point of the circular superhydrophobic surface are all located on the same straight line, as shown in Figure 1b (top view). For this experiment, the centrifugal force of the droplet on the disk can be neglected [33]; therefore, the motion of the surface during droplet impact can be considered as rectilinear. The environmental temperature was about 25 °C, and the humidity was around 30%.

3. Results and Discussion

Figure 2 shows the spreading and receding stages of water droplets on a SHPS at Wen = 29.80, with Wet = 0 and Wet = 44.06 (multimedia view). After the droplet hits the surface, a liquid film forms as it gradually spreads out towards the maximum spreading radius Dmax. The droplet then retracts from both ends towards the center, with its height gradually increasing due to the conversion of surface energy into kinetic energy. The retraction process is dominated by inertia force and surface tension, and it is also influenced by the motion of the surface. Once sufficient kinetic energy has accumulated, the droplet will start to bounce off the surface.
When the surface is stationary, the receding rates of the droplet from both ends are the same (Figure 2a, t = 3.83 ms), resulting in a symmetric bouncing phenomenon. When the surface is in motion, the receding rate on the upstream side is faster than on the downstream side. (Here, downstream refers to the side in the direction of surface movement and the upstream side refers to the opposite direction, as shown in Figure 2b. We will discuss the upstream and downstream in the subsequent sections.) This is due to the influence of surface velocity, causing more kinetic energy to accumulate on the upstream side resulting in a higher height than the downstream side, leading to an “L”-shaped droplet (Figure 2b, t = 5.01~5.83 ms). As the surface velocity increases, the spreading length of the droplet is stretched longer under the influence of the surface forces, making the “L”-shape more pronounced. Subsequently, the droplet begins to bounce off the surface. Compared to the stationary surface, the contact time is reduced from 8.17 ms to 7.17 ms when the surface is in motion. When the surface is stationary, both ends of the droplet simultaneously leave the surface, while in motion, the downstream side of the droplet will detach from the surface before the upstream side (Figure 2b, t = 4.17 ms). This reduces the portion of the droplet in contact with the surface, leading to a reduction in the contact time [34].
In addition, we conducted statistical analysis of experiments involving different Wet and Wen for the contact time. Figure 3a shows the normalized contact time τ/τ0 as a function of Wet under different Wen conditions, where  τ 0 ρ D 0 3 σ  is the inertial-capillary time [35]. As seen, the contact time gradually decreases as Wet increases. The difference between the contact times at Wet = 0 and Wet = 116 is about 30%. These findings are consistent with previous conclusions on moving SHPSs [25,29]. The time from droplet impact surface to maximum spreading is defined as τspread, and the time from maximum spreading to detaching from the surface is defined as τretract. It can be observed that under different Wet, the τspread remains relatively constant, and the variations in τretract and contact time are similar, as shown in Figure 3b.

3.1. Momentum Transfer of Droplet on the Moving Surface

The bouncing and spreading of droplets are essentially the exchange of kinetic energy and surface energy. When a droplet hits a stationary surface, it bounces vertically along the impact direction, while during surface motion the rebound direction of the droplet tilts towards the direction of surface motion. Research has shown that there are two theories explaining the cause of this phenomenon. The first theory, based on the aerodynamic Leidenfrost effect, suggests it is due to the thin layer of air formed between the droplet and the surface [29], while the second theory proposes it is the result of liquid-surface contact forming a viscous boundary layer [31]. We are inclined towards the latter theory. Furthermore, this study is based on the no-slip boundary condition and derives the scaling relationship for the horizontal speed of droplet.
Accordingly, we calculated the center of mass from the shape of droplet in the side view, taking the moment of contact t = 0 ms as the starting point and the moment of rebound as the end point (Figure 4a). By measuring the displacement of the center of mass and the contact time, the averaged horizontal velocity Va of droplet can be obtained.
Initially, Va = 0, but due to the shear effect of the surface the momentum increases during the impact process. Considering the total viscous force and combining Newton’s second law in the horizontal direction mdVa/dt = F(t), the average velocity in the horizontal direction can be obtained [31]:
V a = 3 μ V t 2 ρ D 0 3 1 δ 0 0 t D 2 ( t ) d t
where μ is the viscosity of the liquid, δ0 is the thickness of the boundary layer, t is the contact time of the droplet on the moving surface, and D2(t) represents the contact area of the droplet. In previous studies, D2(t) is considered to be the square of the maximum spreading length on a stationary surface Dmax−s. However, the droplet is not circular on a moving surface. It will be stretched along the direction of surface movement, and the maximum spreading of the droplet in that direction is defined as Dmax−m, as shown in Figure 4c. The maximum spreading length in the direction perpendicular to the horizontal is less influenced by surface motion, so it can be approximated as the spreading length Dmax−s on a stationary surface [36]. Based on this, we can deduce:
V a ~ 3 μ V t 2 ρ D 0 3 1 δ 0 D max m D max s t
With the maximal spreading ratio of the drop βmax = Dmax/D0, the dimensionless integral mean value of boundary layer thickness δ = δ0/D0, the dimensionless contact time τ = Vnt/D0, we can derive that:
V a ~ 3 μ V t 2 ρ D 0 V n 1 δ β max m β max s τ
On the moving surface, the maximum spreading ratio βmax-m~Wen1/4Ca1/6 [29], where the capillary number Ca = μVt/γ. (Despite this equation being based on the aerodynamic Leidenfrost effect, our experimental results have shown good agreement with it; see the comparative results in Supplementary Material Figure S1). On the stationary surface, the maximum spreading ratio βmax-s~Wen1/4(1 − cosθ)−1/2 [37].
Another aspect that distinguishes our work from previous research is that they considered the contact time τ in Equation (3) as the contact time on a stationary surface τs, while we consider τ as the contact time on a moving surface τm. In Figure 3a, we can clearly observe a significant difference in contact time between τs and τm. Based on volume conservation, the relationship between these two is given by [25]:
τ m τ s = D max s D max m
The validation of Equation (3) can be observed in Supplementary Material Figure S2. The contact time on the stationary surface τs~Wen1/2(1 − cosθ)−1/2. By combining Equations (3) and (4) with the above scaling relationships, we can deduce:
V a ~ 3 2 δ ( 1 cos θ ) 5 4 Re 1 We n   Ca 1 12 V t
where the dimensionless Reynolds number is defined as Re = ρVnD0/μ. Equation (5) is derived using the contact time τm and maximum spreading ratio βmax-m on a moving surface. If we were to derive it using the contact time τs and maximum spreading ratio βmax-s on a stationary surface [31], we could derive:
V a ~ 3 2 δ ( 1 cos θ ) 3 2 Re 1 We n V t
When comparing Equations (5) and (6), the difference is the additional influence of the capillary number Ca in Equation (5). This inclusion of Ca enhances the influence of surface motion on the droplet velocity, making it a more accurate representation. The boundary layer thickness can be regarded as a constant value. Therefore, Equation (5) can be expressed as:
V a ~ ( 1 cos θ ) 5 4 Re 1 We n Ca 1 12 V t
In Figure 5, we can observe a good consistency between the experimental data and Equation (7).
As previously mentioned, the conversion of momentum during droplet impact involves the consumption of energy. This implies that the averaged horizontal velocity of the droplet Va should be lower than the surface velocity Vt. We found that the velocity of the droplet is only about 1% of the surface movement velocity. As a result, we defined the averaged restitution coefficient εt = Va/Vt. Previous research has demonstrated a negative correlation between Wen and εt because higher Wen impacts can cause more significant droplet deformation, leading to more energy dissipation during rebound and a decrease in εt [38,39,40,41,42]. However, our study on the moving surface revealed an opposite result; we found that εt increases with the increase in Wen, as shown in Figure 6a, and there is no significant change with the increase in Wet, as shown in Figure 6b. The black line in the box represents the median line, and the point in the middle is the mean value. The upper and lower boundaries of the box represent the upper interquartile and lower interquartile, respectively, while the whiskers represent the extreme values. The reasons for this different trend are as follows. Firstly, previous studies focused on instantaneous velocity, while this study specifically examines the average velocity throughout the entire impact process. Secondly, a significant factor is that as the droplet impacts a moving surface, with an increase in Wen, the spreading length increases by approximately 20%. This results in the viscous boundary layer length between the droplet and the surface also increasing, leading to an increase in the force exerted on the droplet in the horizontal direction. Ultimately, this is manifested as an increase in the coefficient of restitution.

3.2. Morphology of Droplet on Moving Surface

We conducted an analysis of changes in droplet behavior from a momentum perspective. In addition, the most significant difference between the stationary and moving surfaces is not only the contact time, but also the asymmetric spreading and receding of the droplet on the upstream and downstream sides.
Therefore, we divided the droplet into two portions Dup and Ddown based on the impact center point. Initially, the positions of the center points Ds = 0 is represented by the red dot (Figure 7a). It varies at different moments; it moves with the horizontal velocity Va of the droplet, as indicated by the green point in Figure 7a. The distance from the point along the moving direction to the edge of the droplet is referred to as Dup, and the distance from the point along the opposite direction of motion to the edge of the droplet is referred to as Ddown. When Wet = 0, the spreading length of the droplet on the upstream and downstream sides is symmetrically distributed along Ds = 0. However, as Wet increases, the spreading length of the droplet on the upstream side gradually increases, and its spreading time becomes longer, as shown in Figure 8a,b The negative values in Figure 8a indicate that the droplet is located to the right of the boundary point. In addition, the time of initiating receding on the upstream side is later than that on the downstream side; the spreading time of the downstream side is around 2 ms. For different Wet values, there is almost no effect on the length of spreading. By contrast, the length during the receding phase gradually reduces with the increase in Wet. In Figure 8c,d, with a fixed Wet = 44.06 and different Wen, the spreading length on both the upstream and downstream sides increases with Wen, with the increment being more significant on the downstream side. Dup during both the retraction and spreading phases is increasing with Wen, but the increment during the retraction phase is relatively small.
Thus, we postulate that when the droplet reaches its maximum spreading length the length of the upstream and downstream portions should be a function of the impact velocity Vn and surface velocity Vt. Assuming that the averaged spreading velocity is Vs and the surface velocity is Vt, the averaged velocity of the upstream can be treated as Vup = VsVt; similarly, the averaged velocity of the downstream is Vdown = Vs + Vt, as shown in Figure 7c. Taking into account that the droplet will also move a certain distance caused by the moving substrate, the maximal spread length ratio of the upstream to downstream can be expressed as:
D up D down ~ ( V s V t ) τ spread V a τ spread ( V s + V t ) τ spread + V a τ spread
If the droplet’s upstream and downstream are distinguished based on the calculated moving distance of the green dot, then the term Vaτs representing the sliding distance of the droplet in Equation (7) can be eliminated, and the formula can be simplified as:
D up D down ~ V s V t V s + V t
Obviously, Vt~αWet1/2, and for Vs, it is a function of the maximum spreading length and spreading time, Vs~Dmax/τspread. As can be observed in Figure 3b, the spreading time τspread is independent of surface velocity and impact velocity, which is consistent with previous research findings [3,29]. Therefore, we can treat τspread as a constant, while Dmax~D0Wen1/4; based on these, we can derive the expression for Vs~βWen1/4. Substituting these values into Equation (8) we have:
D up D down ~ α We n 1 / 4 β We t 1 / 2 α We n 1 / 4 + β We t 1 / 2
where coefficients α and β are related constants. Equation (10) can be written as Equation (11):
D up D down ~ 1 2 We t 1 / 2 k We n 1 / 4 We t 1 / 2
where coefficient k = α/β. In Figure 9, we can observe a good consistency between the experimental data and Equation (11). This scaling relationship has practical applications such as using water to cool rotating machinery. Understanding the spreading range of droplet holds significance for such applications [43].

4. Conclusions

In this study, experiments have shown that the contact time between water droplets and surfaces is inversely proportional to surface velocity. Our main conclusions are as follows. 1. We have refined the scaling relationship for the horizontal velocity Va of the droplet, substituting the contact time and spreading length of the stationary surface in the scaling relationship with those on a moving surface. These modifications enhance the accuracy of the results. These corrections make the scaling relationship more accurate. 2. Furthermore, an investigation into the variation relationship between the droplet’s tangential restitution coefficient εt and Weber numbers Wen and Wet was conducted. Additionally, we were surprised to discover that the velocity of the droplets was only about 1% of the SHPSs speed. We will focus on exploring ways to improve the exchange rate of droplet momentum in future work. 3. Based on the asymmetry of droplet spreading, a coupling scale relationship between the maximum spreading length ratio of water droplets upstream and downstream was revealed by superimposing the spreading velocity and the surface velocity. This has provided us with new insights into the energy transfer mechanisms of small droplets.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/sym16010029/s1, Figure S1: The validation of the maximum spreading ratio βmax-m on a moving surface; Figure S2: The validation of the contact time τm on a moving surface.

Author Contributions

Writing—original draft, W.W.; Formal analysis, W.W.; Data curation, W.W.; Methodology, W.Y.; Investigation, W.Y.; Software, Z.Y.; Writing—editing, S.C.; Funding acquisition, S.C.; Visualization, D.C.; Supervision, X.L.; Conceptualization, J.Z.; Writing—review & editing, J.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 12002212, 11872283) and Shanghai Sailing Program (Grant No. 20YF1432800). The grant is gratefully acknowledged.

Data Availability Statement

The data that support the findings of this study are available within the article and its Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cheng, X.; Sun, T.-P.; Gordillo, L. Drop impact dynamics: Impact force and stress distributions. Annu. Rev. Fluid Mech. 2022, 54, 57–81. [Google Scholar] [CrossRef]
  2. De Ruiter, J.; Lagraauw, R.; Van Den Ende, D.; Mugele, F. Wettability-independent bouncing on flat surfaces mediated by thin air films. Nat. Phys. 2015, 11, 48–53. [Google Scholar] [CrossRef]
  3. Josserand, C.; Thoroddsen, S.T. Drop impact on a solid surface. Annu. Rev. Fluid Mech. 2016, 48, 365–391. [Google Scholar] [CrossRef]
  4. Enright, R.; Miljkovic, N.; Sprittles, J.; Nolan, K.; Mitchell, R.; Wang, E.N. How coalescing droplets jump. ACS Nano 2014, 8, 10352–10362. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, F.; Ghigliotti, G.; Feng, J.J.; Chen, C.-H. Numerical simulations of self-propelled jumping upon drop coalescence on non-wetting surfaces. J. Fluid Mech. 2014, 752, 39–65. [Google Scholar] [CrossRef]
  6. Liu, X.; Cheng, P.; Quan, X. Lattice Boltzmann simulations for self-propelled jumping of droplets after coalescence on a superhydrophobic surface. Int. J. Heat Mass Transf. 2014, 73, 195–200. [Google Scholar] [CrossRef]
  7. Han, X.; Li, J.; Tang, X.; Li, W.; Zhao, H.; Yang, L.; Wang, L. Droplet bouncing: Fundamentals, regulations, and applications. Small 2022, 18, 2200277. [Google Scholar] [CrossRef]
  8. Yu, X.; Zhang, Y.; Hu, R.; Luo, X. Water droplet bouncing dynamics. Nano Energy 2021, 81, 105647. [Google Scholar] [CrossRef]
  9. Esmaeili, A.R.; Mir, N.; Mohammadi, R. Further step toward a comprehensive understanding of the effect of surfactant additions on altering the impact dynamics of water droplets. Langmuir 2021, 37, 841–851. [Google Scholar] [CrossRef]
  10. Liu, H.; Si, C.; Cai, C.; Zhao, C.; Yin, H. Experimental investigation on impact and spreading dynamics of a single ethanol–water droplet on a heated surface. Chem. Eng. Sci. 2021, 229, 116106. [Google Scholar] [CrossRef]
  11. Fink, V.; Cai, X.; Stroh, A.; Bernard, R.; Kriegseis, J.; Frohnapfel, B.; Marschall, H.; Wörner, M. Drop bouncing by micro-grooves. Int. J. Heat Fluid Flow 2018, 70, 271–278. [Google Scholar] [CrossRef]
  12. Okumura, K.; Chevy, F.; Richard, D.; Quéré, D.; Clanet, C. Water spring: A model for bouncing drops. Europhys. Lett. 2003, 62, 237. [Google Scholar] [CrossRef]
  13. Guo, Y.; Lian, Y.; Sussman, M. Investigation of drop impact on dry and wet surfaces with consideration of surrounding air. Phys. Fluids 2016, 28, 073303. [Google Scholar] [CrossRef]
  14. Stevens, C.S.; Latka, A.; Nagel, S.R. Comparison of splashing in high-and low-viscosity liquids. Phys. Rev. E 2014, 89, 063006. [Google Scholar] [CrossRef] [PubMed]
  15. Zen, T.-S.; Chou, F.-C.; Ma, J.-L. Ethanol drop impact on an inclined moving surface. Int. Commun. Heat Mass Transf. 2010, 37, 1025–1030. [Google Scholar] [CrossRef]
  16. Yu, W.; Zhu, D.; Wang, W.; Yu, Z.; Chen, S.; Zhao, J. The rebounding-coalescing behaviors in drop-on-drop impact on a superhydrophobic surface. Appl. Phys. Lett. 2022, 121, 061602. [Google Scholar] [CrossRef]
  17. Clanet, C.; Béguin, C.; Richard, D.; Quéré, D. Maximal deformation of an impacting drop. J. Fluid Mech. 2004, 517, 199–208. [Google Scholar] [CrossRef]
  18. Kreder, M.J.; Alvarenga, J.; Kim, P.; Aizenberg, J. Design of anti-icing surfaces: Smooth, textured or slippery? Nat. Rev. Mater. 2016, 1, 15003. [Google Scholar] [CrossRef]
  19. Ren, S.; Wang, S.; Dong, Z.; Chen, J.; Li, L. Dynamic behaviors and self-cleaning property of droplet on superhydrophobic coating in uniform DC electric field. Colloids Surf. A Physicochem. Eng. Asp. 2021, 626, 127056. [Google Scholar] [CrossRef]
  20. Bird, J.C.; Dhiman, R.; Kwon, H.-M.; Varanasi, K.K. Reducing the contact time of a bouncing drop. Nature 2013, 503, 385–388. [Google Scholar] [CrossRef]
  21. Liu, Y.; Whyman, G.; Bormashenko, E.; Hao, C.; Wang, Z. Controlling drop bouncing using surfaces with gradient features. Appl. Phys. Lett. 2015, 107, 051604. [Google Scholar] [CrossRef]
  22. Hou, J.; Gong, J.; Wu, X.; Huang, Q. Numerical study on impacting-freezing process of the droplet on a lateral moving cold superhydrophobic surface. Int. J. Heat Mass Transf. 2022, 183, 122044. [Google Scholar] [CrossRef]
  23. Gladkov, S.O. On Some Theoretical Aspects of The Evaporation Process of a Droplet and Its Optimal Size When Extinguishing Fires. Inventions 2023, 8, 35. [Google Scholar] [CrossRef]
  24. Reis, N.; Ainsley, C.; Derby, B. Ink-jet delivery of particle suspensions by piezoelectric droplet ejectors. J. Appl. Phys. 2005, 97, 094903. [Google Scholar] [CrossRef]
  25. Zhang, X.; Zhu, Z.; Zhang, C.; Yang, C. Reduced contact time of a droplet impacting on a moving superhydrophobic surface. Appl. Phys. Lett. 2020, 117, 151602. [Google Scholar] [CrossRef]
  26. Tao, R.; Fang, W.; Wu, J.; Dou, B.; Xu, W.; Zheng, Z.; Li, B.; Wang, Z.; Feng, X.; Hao, C. Rotating Surfaces Promote the Shedding of Droplets. Research 2023, 6, 0023. [Google Scholar] [CrossRef] [PubMed]
  27. Chu, F.; Li, S.; Hu, Z.; Wu, X. Regulation of droplet impacting on superhydrophobic surfaces: Coupled effects of macrostructures, wettability patterns, and surface motion. Appl. Phys. Lett. 2023, 122, 160503. [Google Scholar] [CrossRef]
  28. Gauthier, A.; Bird, J.C.; Clanet, C.; Quéré, D. Aerodynamic Leidenfrost effect. Phys. Rev. Fluids 2016, 1, 084002. [Google Scholar] [CrossRef]
  29. Zhan, H.; Lu, C.; Liu, C.; Wang, Z.; Lv, C.; Liu, Y. Horizontal motion of a superhydrophobic substrate affects the drop bouncing dynamics. Phys. Rev. Lett. 2021, 126, 234503. [Google Scholar] [CrossRef]
  30. García-Geijo, P.; Riboux, G.; Gordillo, J.M. Inclined impact of drops. J. Fluid Mech. 2020, 897, A12. [Google Scholar] [CrossRef]
  31. Wang, M.; Shi, Y.; Wang, S.; Xu, H.; Zhang, H.; Wei, M.; Wang, X.; Peng, W.; Ding, H.; Song, M. Directional droplet bouncing on a moving superhydrophobic surface. Iscience 2023, 26, 106389. [Google Scholar] [CrossRef] [PubMed]
  32. Shu, Y.; Chu, F.; Hu, Z.; Gao, J.; Wu, X.; Dong, Z.; Feng, Y. Superhydrophobic strategy for nature-inspired rotating microfliers: Enhancing spreading, reducing contact time, and weakening impact force of raindrops. ACS Appl. Mater. Interfaces 2022, 14, 57340–57349. [Google Scholar] [CrossRef] [PubMed]
  33. Moqaddam, A.M.; Chikatamarla, S.S.; Karlin, I.V. Drops bouncing off macro-textured superhydrophobic surfaces. J. Fluid Mech. 2017, 824, 866–885. [Google Scholar] [CrossRef]
  34. Moghtadernejad, S.; Jadidi, M.; Hanson, J.; Johnson, Z. Dynamics of droplet impact on a superhydrophobic disk. Phys. Fluids 2022, 34, 062104. [Google Scholar] [CrossRef]
  35. Richard, D.; Clanet, C.; Quéré, D. Contact time of a bouncing drop. Nature 2002, 417, 811. [Google Scholar] [CrossRef]
  36. Almohammadi, H.; Amirfazli, A. Asymmetric spreading of a drop upon impact onto a surface. Langmuir 2017, 33, 5957–5964. [Google Scholar] [CrossRef]
  37. Gao, S.-R.; Wei, B.-J.; Jin, J.-X.; Ye, J.-S.; Wang, Y.-F.; Zheng, S.-F.; Yang, Y.-R.; Wang, X.-D. Contact time of a droplet impacting hydrophobic surfaces. Phys. Fluids 2022, 34, 067104. [Google Scholar] [CrossRef]
  38. Aria, A.I.; Gharib, M. Physicochemical characteristics and droplet impact dynamics of superhydrophobic carbon nanotube arrays. Langmuir 2014, 30, 6780–6790. [Google Scholar] [CrossRef]
  39. Bertola, V. An experimental study of bouncing Leidenfrost drops: Comparison between Newtonian and viscoelastic liquids. Int. J. Heat Mass Transf. 2009, 52, 1786–1793. [Google Scholar] [CrossRef]
  40. Biance, A.-L.; Chevy, F.; Clanet, C.; Lagubeau, G.; Quéré, D. On the elasticity of an inertial liquid shock. J. Fluid Mech. 2006, 554, 47–66. [Google Scholar] [CrossRef]
  41. Chen, L.; Xiao, Z.; Chan, P.C.; Lee, Y.-K.; Li, Z. A comparative study of droplet impact dynamics on a dual-scaled superhydrophobic surface and lotus leaf. Appl. Surf. Sci. 2011, 257, 8857–8863. [Google Scholar] [CrossRef]
  42. Qian, L.; Huo, B.; Chen, Z.; Li, E.; Ding, H. Droplet bouncing on moving superhydrophobic groove surfaces. Int. J. Multiph. Flow 2023, 165, 104454. [Google Scholar] [CrossRef]
  43. Yarin, A.L. Drop impact dynamics: Splashing, spreading, receding, bouncing…. Annu. Rev. Fluid Mech. 2006, 38, 159–192. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic of the experimental setup. (b) View from above showing the position of the camera, droplet impact point, and center of the circular SHPS.
Figure 1. (a) Schematic of the experimental setup. (b) View from above showing the position of the camera, droplet impact point, and center of the circular SHPS.
Symmetry 16 00029 g001
Figure 2. (a) Snapshot of droplet impact on the SHPS for Wen = 29.80 and Wet = 0. (b) Snapshot of droplet impact on the SHPS for Wen = 29.80 and Wet = 44.06. (Multimedia view).
Figure 2. (a) Snapshot of droplet impact on the SHPS for Wen = 29.80 and Wet = 0. (b) Snapshot of droplet impact on the SHPS for Wen = 29.80 and Wet = 44.06. (Multimedia view).
Symmetry 16 00029 g002
Figure 3. (a) Normalized contact time τ/τ0 under different Wen groups varies as the change of Wet, where τ0 is the inertial-capillary time. (b) Spreading time τspread and retraction time τretract vary as the change of Wet.
Figure 3. (a) Normalized contact time τ/τ0 under different Wen groups varies as the change of Wet, where τ0 is the inertial-capillary time. (b) Spreading time τspread and retraction time τretract vary as the change of Wet.
Symmetry 16 00029 g003
Figure 4. (a) The blue dot represents the centroid of the droplet, from which the droplet velocity Va is calculated based on the distance moved in the horizontal direction. (b) Side view of the droplet at different time instances. (c) Top view of the droplet at different time instances, where Dmax-s and Dmax-m denote the maximum spreading lengths of the droplet on stationary and moving surfaces, respectively.
Figure 4. (a) The blue dot represents the centroid of the droplet, from which the droplet velocity Va is calculated based on the distance moved in the horizontal direction. (b) Side view of the droplet at different time instances. (c) Top view of the droplet at different time instances, where Dmax-s and Dmax-m denote the maximum spreading lengths of the droplet on stationary and moving surfaces, respectively.
Symmetry 16 00029 g004
Figure 5. Variation of Va as a function of Re, Wen, Ca, and Vt, with the solid line representing the best fit of Equation (7).
Figure 5. Variation of Va as a function of Re, Wen, Ca, and Vt, with the solid line representing the best fit of Equation (7).
Symmetry 16 00029 g005
Figure 6. (a) The boxplot of tangential recovery coefficient εt with respect to Wen whose range corresponding to each Wet is 5.2–116.1. (b) The boxplot of tangential recovery coefficient εt with respect to Wet whose range corresponding to each Wen is 5.7–60.9.
Figure 6. (a) The boxplot of tangential recovery coefficient εt with respect to Wen whose range corresponding to each Wet is 5.2–116.1. (b) The boxplot of tangential recovery coefficient εt with respect to Wet whose range corresponding to each Wen is 5.7–60.9.
Symmetry 16 00029 g006
Figure 7. (a) Schematic diagram distinguishing the upstream and downstream spreading of the droplet. The red dot represents the initial center of impact of the droplet, while the green dot indicates the position after moving with velocity Va. The green dot serves as the reference for defining the boundary between the upstream and downstream regions, where Ds = 0. (b) Side view of the droplet at different time instances. (c) Schematic representation of the velocity superposition of Vs and Vt.
Figure 7. (a) Schematic diagram distinguishing the upstream and downstream spreading of the droplet. The red dot represents the initial center of impact of the droplet, while the green dot indicates the position after moving with velocity Va. The green dot serves as the reference for defining the boundary between the upstream and downstream regions, where Ds = 0. (b) Side view of the droplet at different time instances. (c) Schematic representation of the velocity superposition of Vs and Vt.
Symmetry 16 00029 g007
Figure 8. (a) Variation of upstream and downstream spread lengths of the droplet with time for different values of Wet at Wen = 49.8, with the center of the droplet at Ds = 0 at the beginning of impact used to distinguish between the upstream and downstream regions. (b) Actual deformation of the droplet corresponding to (a). (c) Variation of upstream and downstream spread lengths of the droplet with time for different values of Wen at Wet = 44.06. (d) Actual deformation of the droplet corresponding to (c).
Figure 8. (a) Variation of upstream and downstream spread lengths of the droplet with time for different values of Wet at Wen = 49.8, with the center of the droplet at Ds = 0 at the beginning of impact used to distinguish between the upstream and downstream regions. (b) Actual deformation of the droplet corresponding to (a). (c) Variation of upstream and downstream spread lengths of the droplet with time for different values of Wen at Wet = 44.06. (d) Actual deformation of the droplet corresponding to (c).
Symmetry 16 00029 g008
Figure 9. Wet and Wen as functions of the normalized ratio of upstream and downstream spread lengths, the solid line representing the best fit of Equation (11).
Figure 9. Wet and Wen as functions of the normalized ratio of upstream and downstream spread lengths, the solid line representing the best fit of Equation (11).
Symmetry 16 00029 g009
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, W.; Yu, W.; Yu, Z.; Chen, S.; Cao, D.; Liu, X.; Zhao, J. Non-Axisymmetric Bouncing Dynamics on a Moving Superhydrophobic Surface. Symmetry 2024, 16, 29. https://doi.org/10.3390/sym16010029

AMA Style

Wang W, Yu W, Yu Z, Chen S, Cao D, Liu X, Zhao J. Non-Axisymmetric Bouncing Dynamics on a Moving Superhydrophobic Surface. Symmetry. 2024; 16(1):29. https://doi.org/10.3390/sym16010029

Chicago/Turabian Style

Wang, Wenhao, Wenlong Yu, Zhiyuan Yu, Shuo Chen, Damin Cao, Xiaohua Liu, and Jiayi Zhao. 2024. "Non-Axisymmetric Bouncing Dynamics on a Moving Superhydrophobic Surface" Symmetry 16, no. 1: 29. https://doi.org/10.3390/sym16010029

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop