Next Article in Journal
Preparation of Memantine-Loaded Chitosan Nanocrystals: In Vitro and Ex Vivo Toxicity Analysis
Previous Article in Journal
Boron Nanotube Structure Explored by Evolutionary Computations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on Preparation of Superhydrophobic and Superoleophobic Surface by Laser Micromachining and Its Hybrid Methods

1
School of Mechanical Engineering, Jiangsu University, Zhenjiang 212013, China
2
College of Mechanical Engineering, Donghua University, Shanghai 201620, China
3
College of Energy and Power, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China
4
Guangdong Provincial Research Center of CNC Technology and Functional Component Engineering Technology, Guangdong Polytechnic Normal University, Guangzhou 510665, China
*
Authors to whom correspondence should be addressed.
Crystals 2023, 13(1), 20; https://doi.org/10.3390/cryst13010020
Submission received: 28 October 2022 / Revised: 26 November 2022 / Accepted: 16 December 2022 / Published: 23 December 2022
(This article belongs to the Section Crystalline Metals and Alloys)

Abstract

:
Functional wetting surfaces have excellent prospects in applications including self-cleaning, anti-fog, anti-icing, corrosion resistance, droplet control, and friction power generation. Laser micromachining technology is an advanced method for preparing such functional surfaces with high efficiency and quality. To fully exploit the potential of laser micromachining and the related hybrid methods, a wide spectrum of knowledge is needed. The present review systematically discusses the process capabilities and research developments of laser micromachining and its hybrid methods considering the research both in basic and practical fields. This paper outlines the relevant literature, summarizes the characteristics of functional wetting surfaces and also the basic scientific requirements for laser micromachining technology. Finally, the challenges and potential applications of superhydrophobic and superoleophobic surface are briefly discussed. This review fills the gap in the research literature by presenting an extended literature source with a wide coverage of recent developments.

1. Introduction

Inspired by plant and animal surfaces in nature, such as self-cleaning lotus leaves and drag-reduction shark skin, the preparation of functional surfaces has received extensive attention [1,2,3]. The research on functional wetting surfaces has become a hotspot due to its outstanding advantages in self-cleaning, anti-icing, anti-corrosion, and other applications [4,5,6,7]. Wettability represents the ability of liquid to maintain contact with solid surface [8,9,10]. The static contact angle (θCA) and dynamic rolling angle (θRA) of the droplets characterize the wettability of the underlying surface [11], as shown Figure 1. In general, when the solid surface θCA is higher than 150° while the θRA is below 10°, the surface is defined as a superhydrophobic surface [12,13,14].
Based on the wetting states of droplets, four typical wettability regimes can be identified [15,16,17]: superhydrophilic, hydrophilic, hydrophobic, and superhydrophobic, as shown Figure 2. Among them, the two extreme phenomena, superhydrophobic and superhydrophilic, have attracted much attention from global researchers, on account of their great application prospects. For example, by imitating the hierarchical structures on a lotus leaf surface, similar microstructures were prepared artificially to obtain self-cleaning ability [18]. Inspired by the unique anisotropic micro-nano structures on a rice leave surface, functional wetting surfaces achieving directional droplet transportation were prepared [19,20,21]. Inspired by the back structure of desert beetles, superhydrophobic surfaces with water collection function are used for water harvesting in arid regions. A water walking robot has been invented by imitating the superhydrophobic structure of the legs of water striders [22].
Wetting properties are determined by surface roughness and chemical composition [23]. A rough surface with low surface energy can easily achieve superhydrophobicity. But typically, superhydrophobic surfaces are usually not oleophobic. However, in the field of industrial production, most parts usually work in the environment of oil–water mixture. Therefore, superamphiphobic surfaces that have both superhydrophobic and superoleophobic abilities are needed. To obtain superoleophobic characteristic, some special surface microstructures such as re-entrant structures need to prepared [24,25]. To obtain superamphiphobic surfaces, a series of methods have been proposed, such as laser micromachining [9,26,27,28,29,30], photolithographing [31], electrochemical machining [32,33], vapor depositing [34,35], anodic oxidation [36], sol–gel [37], plasma treatment [12], sandblasting [38], and self-assembling [39,40].
One of the most concerning methods is laser micromachining technology, which has been widely used in preparing functional surfaces because of its advantages of high efficiency, precision, flexibility, and no pollution. According to the energy output method, the laser can be classified as the continuous laser and the pulsed laser. The pulsed laser can be divided into the short-pulsed laser and the long-pulsed laser according to the pulse duration. The superhydrophobic and superoleophobic surfaces are mainly prepared with short-pulsed lasers (femtosecond, picosecond, and nanosecond laser), which have short action time, high peak power, and low thermal influence. The prepared surface morphology could be easily controlled by the short-pulsed laser [8,18,41]. Laser micromachining technology is mainly based on photothermal effects [26,42,43], as shown in Figure 3. When the laser is transmitted to the surface of the material, the temperature of the area irradiated by the laser rises sharply, which leads part of the material to evaporate and splash into the air. When these high-temperature splashes encounter air, they are re-cooled to realize re-condensation. Finally, the formation of the microstructure is the result of the splatter accumulation and laser removal. The micro-nanoscale surface structures produced by laser direct processing can be divided into laser-induced periodic structures, micro-nano arrays, and various irregular structures. As shown in Figure 4, with the development of laser micromachining, functional surfaces have been widely used in various industrial fields, such as self-cleaning [44,45], drag reduction [46,47], droplet control [48,49], oil–water separation [50,51], and anti-icing [31,52,53,54,55]. In addition, to further improve the wettability of functional surfaces, various laser hybrid processing technologies incorporating a variety of other technologies are also developing vigorously [56,57,58].
In this paper, the machining methods and the latest progress of superhydrophobic or superamphiphobic surfaces prepared by laser micromachining and its hybrid methods are systematically reviewed. Firstly, the classical wetting model and biomimetic structure are introduced. Then, the preparation strategy and characteristics of functional wetted surfaces are reviewed. Then, the development of laser machining and laser hybrid methods are emphatically introduced. Finally, the challenges and prospects of superamphiphobic surfaces fabricated by laser and its hybrid methods are briefly discussed.

2. Wetting Model and Biological Background

2.1. Wetting Model

To better understand the functional wetting surface, three wetting models are briefly described [1,59,60,61,62]. First, Young’s equation was proposed to describe the wetting characteristics of droplets on an ideal smooth solid surface [63]. Where γ L G , γ S G and γ S L represent liquid gas, solid gas, and solid liquid interfacial tensions, respectively (Figure 5a), θ is the contact angle (CA) of a droplet on the smooth surface.
γ L G c o s θ = γ S G γ S L
Wenzel [64] developed Young’s equation and proposed a theoretical model (Formula (2)) to describe the CA of rough surfaces. Where θw is the apparent CA, r is the ratio of the actual surface area of the rough surface to its projected surface area on the horizontal plane.
c o s θ W = r ( γ S G γ S L / γ L G ) = r c o s θ
In this model, droplets completely wet the pits and grooves on the solid surface, as shown in Figure 5b. It can be concluded that the roughness can increase wetting and anti-wetting, compared to flat surfaces. When the solid surface is hydrophobic (oily), the surface roughness will increase the CA. When the solid surface is hydrophilic (oily), the increase in surface roughness will lead to the decrease in CA. The Wenzel model can only be applicable to smooth interfaces while it cannot be used for non-smooth surfaces. Therefore, Cassie and Baxter proposed the Cassie–Baxter [65] model (Figure 5c). As the droplet is not fully wetted, so the droplet cannot completely fill the micro-pits or micro-grooves. Part of the space of the micro-pits or micro-grooves are occupied by air which will suspend the droplets. The contact state is a composite contact of the solid–liquid–gas. The Cassie–Baxter equation is as follows. Where θC-B is the actual CA, f1 and f2 is the area fraction of solid–liquid contact and gas–liquid contact, respectively.
c o s θ C B = f 1 c o s θ f 2 = f 1 ( 1 + c o s θ ) 1
In addition to the above three most typical models, there are more advanced models to predict the specific wetting state, which are mentioned in the literature [66,67,68].

2.2. Biological Background

2.2.1. Superhydrophobic Surfaces in Nature

For a long time, creatures in nature have evolved to adapt to environmental changes. The wetting characteristics of animal and plant surfaces depend on the structures and chemistry, such as lotus leaves [69,70,71,72], rice leaves [73,74,75], water strider whole body [76,77,78,79], and rose petals [5,80,81,82,83,84]. Barthlott and Neinhuis found the reason why water droplets could roll freely on a lotus leaf surface was that its surface microstructure and waxy layer could hold up water droplets [72]. Jiang et al. [85] proposed that the papillary micron-structures and nanostructures of different shapes combined to form a composite structure. The CA of water droplets on the lotus leaf surface is more than 160° while the RA is less than 5°, which indicates the lotus leaf surface has a perfect superhydrophobic performance. As shown in Figure 6a,b, water droplets are spherical on a lotus leaf surface and they can easily roll off the surface. Rice leaves show superhydrophobic properties, and water droplets roll on the surface showing anisotropy. Guo et al. [86] found that the rice surfaces also have a binary structure (Figure 6c,d). The vein trend of the microstructure is parallel to the rice leaf tip, which is different from the lotus leaf surface microstructure, leading to the anisotropy of the water droplets. Feng et al. [87] studied a special re-entrant structure with a corrugated structure surrounding. As shown in Figure 6e,f, the special structure makes the water droplets firmly adhere to the surface of the petal. The water drop CA is equivalent to that on the lotus leaf surface. The water drop cannot roll off the surface, indicating that the RA is also large [80]. The water strider is an insect that can move freely on water without adhering to it (Figure 6g). Jiang et al. [79] found that the special ability of the water strider was derived from the hierarchical microstructure of the legs. As shown in Figure 6h, there are dense micron scale needle structures facing the same direction on the water strider legs, and each needle structure is arranged with dense nano-groove structures. A composite structure was formed through the effective combination of the micron needle and the complex nano-grooves, which can store air and then make the surface form a stable Cassie state. Butterfly wings are also a typical anisotropic superhydrophobic surface. Zheng et al. [82] observed the surface structure of butterfly wings in detail, and found that the surface of butterfly wings is covered with stacked rectangular scales, and a large number of stripes are arranged orderly on the stacked rectangular scales. As shown in Figure 6i,j, the micron-stripe, nano-stripe, and nano-sheet on the surface will separate from each other when the butterfly wings flap downward. In this case, the gaps between the surface structures resulted in the formation of discontinuous gas–liquid–solid contact between the water droplets and the surface structures. The water droplets RA of the surface of the butterfly wings along the direction of the grooved stripes is small, showing excellent superhydrophobicity. The mosquito compound eye is composed of a large number of micro-sized eyelets, and each eyelet has a nano-sized convex structure [88], as shown in Figure 6k,l. The low surface energy makes the mosquito compound eyes have a superhydrophobic property. Even in the high humidity rain and fog environment, the mosquito compound eyes still remain dry.

2.2.2. Superamphiphobic Surfaces in Nature

In nature, the surfaces of some special animals or plants both have superhydrophobic characteristic and superoleophobic characteristic. This special characteristic is called superamphiphobic. To obtain superamphiphobic characteristic, research have been conducted on the natural organisms which living in humid environment for a long time [23]. As shown in Figure 7a, the skin of the collembola, which often lives in sewage, has excellent superamphiphobic porosity, which can prevent various organic liquids from wetting the skin. The scanning electron microscope (SEM) images show that the skin of the collembola is covered with a large number of nano-particle diamond grids and bristles. The re-entrant microstructure promotes the oil resistance of the surface skin of collembola [23,89,90]. Sun et al. [91] reported the superamphiphobic capacity of various cicada skin, as shown in Figure 7b. A special kind of fiber particles are evenly distributed on the surfaces of the cicada, as shown in Figure 7c. These particles present a hollow spherical structure with a diameter of about 200~700 nm, and the overall morphology is honeycomb. These structures have special morphology, which is called re-entrant curvatures. These special morphologies are conducive to achieving superamphiphobic [92].
In conclusion, organisms with wetting properties in nature have provided researchers with a wealth of ideas for preparing special wetting surfaces. Inspired by these organisms, it has been found that rough surfaces and low surface energy are necessary conditions for achieving superhydrophobic, while special surface re-entrant structures with low surface energy are important features for obtaining superamphiphobic. Based on these findings, researchers have prepared various superhydrophobic surfaces or superhydrophobic surfaces by various methods. Among those methods, laser micromachining technology is promising, because of its high precision and strong controllability [19,56,93,94].

3. Laser Micromachining and Its Hybrid Variants

As an advanced micro-nano manufacturing technology, laser micromachining technology has unique advantages. It can achieve precise processing of almost all kinds of materials with high processing flexibility [52,95,96]. In addition, it is a green processing method, which does not produce waste liquid or other pollution. Based on the above advantages, the laser micromachining technology is an ideal method for micro-nano structure processing. Through laser micromachining technology and its hybrid methods, researchers have prepared various types of functionally wetted surfaces [97,98].

3.1. Laser Micromachining

The laser micromachining for preparing functional wetting surfaces mainly includes the following processes. Firstly, laser scanning the sample surface in a single direction or in a network way to generate regular micron array structure. Secondly, chemical modification is carried out to generate abundant random nanostructures on regular micron array obtained from previous processing. Through the combined processing of the above two steps, a multilayer micro-nano structure is obtained.
Long et al. [99] prepared the microstructure by laser ablating the copper substrate, as shown in Figure 8a. With the change of scanning speed of laser processing parameters, the rolling CA of the prepared sample changed from 0° to 90°, the contact state of water droplets changed from the Cassie state to the Wenzel state, and the surface functionality changed from the self-cleaning to the highly adherent superhydrophobic surface. Hauschwitz et al. [100] obtained three structures in Figure 8b by nanosecond laser scanning. The sample was placed in a vacuum environment after laser processing, which accelerated the chemical adsorption of the organic molecules. The time required for the transition from hydrophilic state to hydrophobic state was reduced to hours, and the sample had perfect superhydrophobicity with a CA of 170° and a RA of 5°. Rajab et al. [101] produced 24 types of superhydrophobic structures on the surfaces of stainless steel using picosecond and nanosecond lasers. The stability of different superhydrophobic surfaces under room-temperature, high-temperature, and ice-water environments were investigated. It was found that the surfaces with large contact angles, high surface roughness, and high fluorine could keep drying underwater longer than other surfaces (Figure 8c). To obtain a superhydrophobic surface of stainless steel with good anti-ice performance, Gaddam et al. [102] applied lubricant to the periodic structure surface induced by laser. Huerta Murillo et al. [103] used a femtosecond laser to induce periodic surface structure on the array micro-column of the titanium alloy surface processed by a nanosecond laser, which has excellent superhydrophobic property when modified by polyethylene gas.
In addition, laser micromachining technology can be applied to the processing of a variety of materials [104,105]. Liu et al. [106] prepared micro/nanostructures on polydimethylsiloxane directly to achieve a superhydrophobic surface using a picosecond laser (Figure 9a). Lee et al. [107] processed micro/nanostructures on the endoscope lens to obtain the anti-fog and anti-biofouling capability. Moreover, those micro/nanostructures improved the transparency and hydrophobic properties of the endoscope lens (Figure 9b). To obtain a lens surface with good transparency and antifouling ability different from Lee et al., Karkantonis et al. [108] firstly induced the periodic structure on the stainless steel by laser, and then copied the generated periodic structure to the plastic sheet for optical lens by hot embossing.
Fan et al. [109] created a stable superamphiphobic surface without other auxiliary processes by the femtosecond laser micromachining of a micro-nano composite structure on the surface of the polytetrafluoroethylene (PTFE) (Figure 10a). The micro-nano composite structure on the surface of the PTFE increased the surface roughness, which made the oil droplet CA reached 153.35°. Lian et al. [97] realized the preparation of the superamphiphobic aluminum alloy surface by using the nanosecond laser to ablate the reticular micro texture in one step. As shown in Figure 10b, the superamphiphobic properties of the microstructure surface were obtained by high-temperature treatment and fluorosilane treatment. Through the study of the oil repellency and corrosion resistance of the two substrates, it found that the surface of the aluminum alloy sample had a better superamphiphobic porosity after fluorination when the scanning interval was 20 μm. The dynamic potentio-dynamic polarization test showed that the superamphiphobic surface had better corrosion resistance than the superhydrophobic surface, which could provide effective protection for bare aluminum alloy. After the prepared surfaces were tested in a large pH range, it was found that the surface with superamphiphobic ability also had good chemical stability compared with other surfaces. Samanta et al. [110] used nanosecond pulse laser surface texture technology to prepare microgrooves on the surface of aluminum alloy 6061. The textured surface was soaked in several chemical solutions to make the target functional groups adhere to the surface to achieve the ultimate extreme wettability. Fluorinated groups (-CF2- and -CF3) with extremely low dispersive and non-dispersive surface energy were introduced to achieve superamphiphobic, thus promoting the surface to repel water and diiodomethane (Figure 10c). Ma et al. [111] prepared a superamphiphobic surface by the laser etching of the network structure on the stainless-steel surface and the modification of low surface energy solution (Figure 10d). The researchers found that the CAs of water, glycerin, peanut oil and hexadecyl ether all exceeded 150° (165.77°, 155.09°, 155.24°, and 152.83°, respectively), while the SAs of water, glycerin, peanut oil and hexadecyl ether were 1°, 2°, 4°, and 10°, respectively. After the self-cleaning durability and wear resistance tests, it was found that the sample had good stability. After three meters of wear loss, the surface of the sample still had superoleophobic capacity, showing good wear resistance. After being placed under natural conditions for 60 days, all sample surfaces still maintained superamphiphobic properties.
In summary, laser micromachining technology has substantially reduced the difficulty of preparing superhydrophobic surfaces and other wetting structures, and it has been a simple and feasible process method for achieving superhydrophobic surfaces. However, the stability and durability of laser-prepared micro-array structures are weak because that they cannot withstand the damage and impact of complex and changeable environment, so their industrial application prospects are limited. To prepare the superhydrophobic surfaces with perfect stability and durability, laser hybrid micromachining technology is proposed and studied.

3.2. Laser Hybrid Micromachining

Superamphiphobic properties can be obtained after lowering the surface energy by laser micromachining technology, but the prepared superhydrophobic structure is prone to gradually lose its inherent characteristics under the external impact. To prepare more stable superamphiphobic structures with impact resistance, laser hybrid micromachining technology is proposed and becomes a research hotspot. A variety of laser hybrid micromachining technologies emerge endlessly, such as laser-chemical etching [112,113], laser-ion implantation [114,115], laser oxidation [116,117,118], laser deposition [33,98,119,120], and laser coating [53,121]. As the methods of laser oxidation and laser deposition can significantly improve the stability of the prepared superamphiphobic surface, this hybrid manufacturing technology is mainly introduced in this manuscript.
Wang et al. [118] used anodic oxidation and laser technology to make superamphiphobic surfaces on titanium. Firstly, laser technology was applied to form micro-convex-array textures on the surface of titanium. Then, the sample was oxidized in the electrolyte of ethylene glycol, 3% hydrofluoric acid and 0.5 wt% NH4F to obtain the TiO2 nanotube arrays. Finally, a perfect superamphiphobic surface was obtained after fluorination. The CAs of water and oil both exceeded 160° while the RAs were both less than 3°. The reversible change of oil droplet adhesion without losing the superoleophobic characteristic was achieved. Magnets were used to control magnetic oil droplets, which could adsorb dust particles on the surface, showing the application potential of self-cleaning.
As shown in Figure 11a, Abele et al. [122] used a pulsed laser to texture the surface of aluminum alloy. SiO2 micro-spheres with uneven distribution were deposited on the aluminum surface by electrophoretic deposition, and then the microstructures with controllable surface roughness were formed. This was also known as the combination of layered structures, which greatly meets the requirements of hierarchical surface morphology. Finally, the superamphiphobic property was obtained after the sample surface was fluorinated. The CA of diiodomethane on the sample surface was 157 ± 1°, which indicated that the prepared sample surface had excellent superamphiphobic sparsity.
To further improve the performance of superamphiphobic surface, the additive–subtractive hybrid manufacturing method was proposed. Han et al. [123] combined laser ablation and chemical bath processing to prepare a layered surface with re-entrant nano CuO on a copper micro-cone. The specific process flow is shown in Figure 11b. Firstly, the micro-cone structure was ablated by laser. Then, the copper plate with the microsphere structure was immersed in a mixed solution of 0.13 mol·L−1 (NH4)2S2O4 and 2.5 mol·L−1 NaOH at room temperature for 20 min. The growth of the nanostructures is greatly promoted on the surface of the sample. On the other hand, many upward convex nanostructures become laterally grown nanostructures on the surface of the micro-cone, which significantly enriched the re-entrant geometry. The layered surface showed outstanding superamphiphobicity with low surface tension. This is because that the fluorosilane can reduce the surface energy and nano-grass structures create a large number of re-entrant geometries. The durability and stability of the superamphiphobic surface were also analyzed. The results showed that superamphiphobic surfaces exhibit excellent long-term durability and high-temperature durability after being placed in air for 7 months, or in water or ethanol for 10 days, or tempered in air at 200 °C. The superamphiphobic surface could withstand 70 tape peeling cycles, 5 wear cycles, 9 min solid particle impact or 20 min water mist impact. The water or cetane droplet CA exceed 150° with the sliding angle at about 20°, which indicated that the prepared surface had perfect comprehensive mechanical durability. The superamphiphobic copper sheet was successfully prepared by femtosecond laser ablation combined with chemical etching. The results showed that the realization of excellent superhydrophobicity was inseparable from the ordered microstructures and the re-entrant nanostructures. Electrodeposition could prepare controllable nanostructures by controlling electrodeposition parameters, which would change the deposition effects [124,125]. Finally, different deposition states and properties were obtained.
After analyzing the research progress in the preparation of superamphiphobic surfaces, Wang et al. [98] discussed the advantages and disadvantages of the current design strategy of the superamphiphobic re-entrant structures. Combined with the surface wetting theory and the existing problems, a new laser-electrochemical-combination process was used to prepare the re-entrant structures to achieve superamphiphobic. This method mainly consists of three parts, including picosecond laser ablation, electropolishing and electrodeposition. The detailed process is as follows. Firstly, picosecond laser scanning along the network cross line is used to obtain the regular distribution of micro-copper-cone structure. Secondly, the micro copper-cone obtained by laser ablation was electropolished. Thirdly, the rich-nickel-cone structure is deposited on the polished micro-copper-cone surface by twice electrodeposition. Finally, low surface energy materials were used to modify the surface microstructure. The measurement results showed that the static CA of the water and rapeseed oil drops on the sample surface prepared by this process was greater than 150°, while the sliding angle SA was less than 10°. In the low temperature test, the droplets have good durability and the ability to resist gravity movement along the inclined plane (Figure 11c).
Song et al. [120] opened up a new way for the preparation of re-entrant structures and the acquisition of superamphiphobic surfaces. They fluorinated the surface of the sample after the laser ablation of the micro-cone array, and then applied electrodeposition technology to prepare the micro-roof array on the top of the micro-column array to form the micro-re-entrant structure. Finally, a further fluorination treatment was carried out to further reduce the surface energy, which realized the superamphiphobic of water and peanut oil. Using a CuSO4 aqueous solution as the electrolyte, copper as the anode, material as the cathode, the micro-roof array was prepared on the top of the micro-column array by electrochemical deposition. In the fabrication of the superhydrophobic micro-column arrays by nanosecond laser ablation, the electrolyte droplets were located at the top of the micro-column and could not be immersed in the interior of the micro-column, so the shape of the liquid/gas interface could be adjusted. After the sample surface was fluorinated again, the water CA reached 158° while the peanut oil CA reached 156° (Figure 11d).
Different from the previous hybrid processing strategy of subtractive manufacturing before additive manufacturing, some scholars proposed the method of preparing the superamphiphobic surface by multiple-step additive manufacturing [126,127,128]. Hokkanen et al. [129] used laser additive, atomic layer deposition and fluoride treatment to obtain surfaces with superior performance. First of all, to obtain the dual re-entrant micro-columns, laser direct writing (DLW) was used to prepare micro-columns and special nanostructures. Then, 50 nm Al2O3 was coated on the sample surface by atomic layer deposition (ALD). The sample surface was then fluorinated with perfluorooctyl trichlorosilane, and the lubricating oil was immersed in the top of the micro-column with the tip of the capillary. Through the micro-nano structure at the top, smooth lubricant was prevented from diffusing to the stable lubricating film. The super lubricating surface was combined with the superoleophobic surface. The composite surface prepared by multiple-step additive manufacturing was more stable than the conventional lubricating surface or the superoleophobic surface, showing advanced anti-icing and anti-fouling capabilities.
As shown in Table 1, compared with a single manufacturing process, laser hybrid micromachining technologies not only can improve the stability of the superamphiphobic surfaces by producing dense micro-nano structures, but also extend the service life of the prepared surface.

4. Challenges and Development Direction

Because of their excellent wetting performance, the superhydrophobic and superoleophobic surfaces have been widely used in self-cleaning, anti-fog, anti-icing, corrosion resistance, liquid nondestructive transfer, oil–water separation, friction power generation, chip laboratory, liquid drop sensor, etc. [48,130,131,132,133]. However, the manufacture and application of those surfaces with special wetting abilities are still in the experimental research stage. There are many problems, from laboratory research to industrial applications.
(1) Currently, laser micromachining technology has greatly improved the preparation efficiency of superhydrophobic and superoleophobic surfaces. However, there is still the problem of high preparation costs for large area preparation.
(2) The natural modification time for the prepared surface without infiltration is too long. The efficiency needs to be improved, while the use of low surface-energy-solution modification is expensive and environmentally polluting. Therefore, laser hybrid processing technology, which can improve processing efficiency and reduce manufacturing costs, is the trend of future technology development.
(3) Superamphiphobic surfaces are susceptible to mechanical damage such as friction and impact, which may cause them to lose their properties. Laser-deposition technology and laser-coating technology can solve this problem and enhance the stability of the surface by producing dense micro-nano structures. With the continuous development of laser-deposition technology and laser-coating technology, the superamphiphobic surface with ultra-high stability and the ability to resist external shocks will be obtained.
(4) To expand the practical application of superhydrophobic and superoleophobic surfaces, the research on the application in the industry should be enhanced. The primary research trend in this field is to endow the superhydrophobic and superoleophobic surface with multiple functions. With endless scientific research institutions and researchers committed to realizing a powerful superhydrophobic and superoleophobic surface, the enormous commercial value and potential would be exploited in the future.

Author Contributions

Conceptualization, Y.L. and M.W.; methodology, Z.Z.; validation, Y.L. and Z.Z.; formal analysis, H.Z.; investigation, H.L.; resources, Z.W.; data curation, Y.W.; writing—original draft preparation, Y.L.; writing—review and editing, M.W.; visualization, D.Z.; supervision, C.G.; project administration, Y.L.; funding acquisition, Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Science and Technology Innovation Major Project of Wenzhou (No. ZG2022007), Natural Science Foundation of Jiangsu Province (No. BK20210755), China Postdoc-toral Science Foundation (No. 2022M710061), Natural Science Foundation of Zhejiang Province (No. LD22E050001) and Natural Science Research of Jiangsu Higher Education Institutions of China (No. 21KJB460014).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the reviewers and the editors for the improvement and the publication of this paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, Y.-L.; Xia, H.; Kim, E.; Sun, H.-B. Recent developments in superhydrophobic surfaces with unique structural and functional properties. Soft Matter 2012, 8, 11217–11231. [Google Scholar] [CrossRef]
  2. Ghaffari, S.; Aliofkhazraei, M.; Darband, G.B.; Zakeri, A.; Ahmadi, E. Review of superoleophobic surfaces: Evaluation, fabrication methods, and industrial applications. Surf. Interfaces 2019, 17, 100340. [Google Scholar] [CrossRef]
  3. Chen, J.; Guo, H.; Zheng, J.; Huang, Y.; Liu, G.; Hu, C.; Wang, Z.L. Self-Powered Triboelectric Micro Liquid/Gas Flow Sensor for Microfluidics. ACS Nano 2016, 10, 8104–8112. [Google Scholar] [CrossRef] [PubMed]
  4. Si, Y.; Dong, Z.; Jiang, L. Bioinspired Designs of Superhydrophobic and Superhydrophilic Materials. ACS Central Sci. 2018, 4, 1102–1112. [Google Scholar] [CrossRef] [PubMed]
  5. Xu, S.; Wang, Q.; Wang, N. Chemical Fabrication Strategies for Achieving Bioinspired Superhydrophobic Surfaces with Micro and Nanostructures: A Review. Adv. Eng. Mater. 2021, 23, 2001083. [Google Scholar] [CrossRef]
  6. Zhu, Y.; Yang, F.; Guo, Z. Bioinspired surfaces with special micro-structures and wettability for drag reduction: Which surface design will be a better choice? Nanoscale 2021, 13, 3463–3482. [Google Scholar] [CrossRef]
  7. Habib, A.; Wu, S.; Fan, Q.; Magu, T.O.; Yao, X.; Lv, J.; Wang, J. Bioinspired in situ repeatable self-recovery of superhydrophobicity by self-reconstructing the hierarchical surface structure. Chem. Commun. 2021, 57, 8425–8428. [Google Scholar] [CrossRef]
  8. Ijaola, A.O.; Bamidele, E.A.; Akisin, C.J.; Bello, I.T.; Oyatobo, A.T.; Abdulkareem, A.; Farayibi, P.K.; Asmatulu, E. Wettability Transition for Laser Textured Surfaces: A Comprehensive Review. Surf. Interfaces 2020, 21, 100802. [Google Scholar] [CrossRef]
  9. Long, J.; Fan, P.; Gong, D.; Jiang, D.; Zhang, H.; Li, L.; Zhong, M. Superhydrophobic Surfaces Fabricated by Femtosecond Laser with Tunable Water Adhesion: From Lotus Leaf to Rose Petal. ACS Appl. Mater. Interfaces 2015, 7, 9858–9865. [Google Scholar] [CrossRef]
  10. Piscitelli, F.; Tescione, F.; Mazzola, L.; Bruno, G.; Lavorgna, M. On a simplified method to produce hydrophobic coatings for aeronautical applications. Appl. Surf. Sci. 2019, 472, 71–81. [Google Scholar] [CrossRef]
  11. Gu, Y.; Zhang, W.; Mou, J.; Zheng, S.; Jiang, L.; Sun, Z.; Wang, E. Research progress of biomimetic superhydrophobic surface characteristics, fabrication, and application. Adv. Mech. Eng. 2017, 9, 1–13. [Google Scholar] [CrossRef]
  12. Das, S.; Kumar, S.; Samal, S.K.; Mohanty, S.; Nayak, S.K. A Review on Superhydrophobic Polymer Nanocoatings: Recent Development and Applications. Ind. Eng. Chem. Res. 2018, 57, 2727–2745. [Google Scholar] [CrossRef]
  13. Dimitrakellis, P.; Gogolides, E. Hydrophobic and superhydrophobic surfaces fabricated using atmospheric pressure cold plasma technology: A review. Adv. Colloid Interface Sci. 2018, 254, 1–21. [Google Scholar] [CrossRef]
  14. Yang, C.-J.; Mei, X.-S.; Tian, Y.-L.; Zhang, D.-W.; Li, Y.; Liu, X.-P. Modification of wettability property of titanium by laser texturing. Int. J. Adv. Manuf. Technol. 2016, 87, 1663–1670. [Google Scholar] [CrossRef] [Green Version]
  15. Durand, N.; Mariot, D.; Améduri, B.; Boutevin, B.; Ganachaud, F. Tailored Covalent Grafting of Hexafluoropropylene Oxide Oligomers onto Silica Nanoparticles: Toward Thermally Stable, Hydrophobic, and Oleophobic Nanocomposites. Langmuir 2011, 27, 4057–4067. [Google Scholar] [CrossRef] [PubMed]
  16. Li, F.; Wang, Z.; Huang, S.; Pan, Y.; Zhao, X. Flexible, Durable, and Unconditioned Superoleophobic/Superhydrophilic Surfaces for Controllable Transport and Oil-Water Separation. Adv. Funct. Mater. 2018, 28, 1706867. [Google Scholar] [CrossRef]
  17. Dong, J.; Wang, Q.; Zhang, Y.; Zhu, Z.; Xu, X.; Zhang, J.; Wang, A. Colorful Superamphiphobic Coatings with Low Sliding Angles and High Durability Based on Natural Nanorods. ACS Appl. Mater. Interfaces 2016, 9, 1941–1952. [Google Scholar] [CrossRef]
  18. Zhang, M.; Feng, S.; Wang, L.; Zheng, Y. Lotus effect in wetting and self-cleaning. Biotribology 2016, 5, 31–43. [Google Scholar] [CrossRef]
  19. Vorobyev, A.Y.; Guo, C. Direct femtosecond laser surface nano/microstructuring and its applications. Laser Photonics Rev. 2013, 7, 385–407. [Google Scholar] [CrossRef]
  20. Yang, C.; Yang, K.; Li, M.; Chen, F.; Yang, Z. The investigation of droplet directional self-transport ability on the slippery liquid-infused surface with anisotropic structure. Prog. Org. Coat. 2022, 168, 106857. [Google Scholar] [CrossRef]
  21. Choi, J.; Park, S.; Jeon, S.; Yim, C. A Facile and Rapid Fabrication of Multifunctional Color-Tuned Titanium. Crystals 2020, 10, 1065. [Google Scholar] [CrossRef]
  22. Zhang, Y.; Zhang, Z.; Yang, J.; Yue, Y.; Zhang, H. A Review of Recent Advances in Superhydrophobic Surfaces and Their Applications in Drag Reduction and Heat Transfer. Nanomaterials 2021, 12, 44. [Google Scholar] [CrossRef]
  23. Rajab, F.H.; Liauw, C.M.; Benson, P.S.; Li, L.; Whitehead, K.A. Production of hybrid macro/micro/nano surface structures on Ti6Al4V surfaces by picosecond laser surface texturing and their antifouling characteristics. Colloids Surf. B Biointerfaces 2017, 160, 688–696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Yong, J.; Chen, F.; Yang, Q.; Huo, J.; Hou, X.J. Superoleophobic surfaces. Chem. Soc. Rev. 2017, 46, 4168–4217. [Google Scholar] [CrossRef] [Green Version]
  25. Liu, M.; Wang, S.; Wei, Z.; Song, Y.; Jiang, L. Bioinspired design of a superoleophobic and low adhesive water/solid interface. Adv. Mater. 2009, 21, 665–669. [Google Scholar] [CrossRef]
  26. Rukosuyev, M.V.; Lee, J.; Cho, S.J.; Lim, G.; Jun, M.B. One-step fabrication of superhydrophobic hierarchical structures by femtosecond laser ablation. Appl. Surf. Sci. 2014, 313, 411–417. [Google Scholar] [CrossRef]
  27. Martínez-Calderon, M.; Rodríguez, A.; Dias-Ponte, A.; Morant-Miñana, M.; Gómez-Aranzadi, M.; Olaizola, S. Femtosecond laser fabrication of highly hydrophobic stainless steel surface with hierarchical structures fabricated by combining ordered microstructures and LIPSS. Appl. Surf. Sci. 2016, 374, 81–89. [Google Scholar] [CrossRef]
  28. Jagdheesh, R.; García-Ballesteros, J.; Ocaña, J. One-step fabrication of near superhydrophobic aluminum surface by nanosecond laser ablation. Appl. Surf. Sci. 2016, 374, 2–11. [Google Scholar] [CrossRef]
  29. Yang, Z.; Liu, X.; Tian, Y. Insights into the wettability transition of nanosecond laser ablated surface under ambient air exposure. J. Colloid Interface Sci. 2019, 533, 268–277. [Google Scholar] [CrossRef] [Green Version]
  30. Dinh, T.-H.; Ngo, C.-V.; Chun, D.-M. Controlling the Wetting Properties of Superhydrophobic Titanium Surface Fabricated by UV Nanosecond-Pulsed Laser and Heat Treatment. Nanomaterials 2018, 8, 766. [Google Scholar] [CrossRef]
  31. Liu, R.; Chi, Z.; Cao, L.; Weng, Z.; Wang, L.; Li, L.; Saeed, S.; Lian, Z.; Wang, Z. Fabrication of biomimetic superhydrophobic and anti-icing Ti6Al4V alloy surfaces by direct laser interference lithography and hydrothermal treatment. Appl. Surf. Sci. 2020, 534, 147576. [Google Scholar] [CrossRef]
  32. Shen, L.; Fan, M.; Qiu, M.; Jiang, W.; Wang, Z. Superhydrophobic nickel coating fabricated by scanning electrodeposition. Appl. Surf. Sci. 2019, 483, 706–712. [Google Scholar] [CrossRef]
  33. Wang, Z.; Shen, L.; Jiang, W.; Fan, M.; Liu, D.; Zhao, J. Superhydrophobic nickel coatings fabricated by scanning electrodeposition on stainless steel formed by selective laser melting. Surf. Coat. Technol. 2019, 377, 124886. [Google Scholar] [CrossRef]
  34. Zheng, J.; Yang, J.; Cao, W.; Huang, Y.; Zhou, Z.; Huang, Y.-X. Fabrication of transparent wear-resistant superhydrophobic SiO2 film via phase separation and chemical vapor deposition methods. Ceram. Int. 2022, 48, 32143–32151. [Google Scholar] [CrossRef]
  35. Fu, J.; Sun, Y.; Ji, Y.; Zhang, J. Fabrication of robust ceramic based superhydrophobic coating on aluminum substrate via plasma electrolytic oxidation and chemical vapor deposition methods. J. Mater. Process. Technol. 2022, 306, 117641. [Google Scholar] [CrossRef]
  36. Li, S.-Y.; Xiang, X.-G.; Ma, B.-H.; Meng, X.-D. Facile preparation of diverse alumina surface structures by anodization and superhydrophobic surfaces with tunable water droplet adhesion. J. Alloys Compd. 2019, 779, 219–228. [Google Scholar] [CrossRef]
  37. Latthe, S.S.; Terashima, C.; Nakata, K.; Sakai, M.; Fujishima, A. Development of sol–gel processed semi-transparent and self-cleaning superhydrophobic coatings. J. Mater. Chem. A 2014, 2, 5548–5553. [Google Scholar] [CrossRef]
  38. Sun, R.; Zhao, J.; Li, Z.; Mo, J.; Pan, Y.; Luo, D. Preparation of mechanically durable superhydrophobic aluminum surface by sandblasting and chemical modification. Prog. Org. Coat. 2019, 133, 77–84. [Google Scholar] [CrossRef]
  39. Celik, N.; Torun, I.; Ruzi, M.; Esidir, A.; Onses, M.S. Fabrication of robust superhydrophobic surfaces by one-step spray coating: Evaporation driven self-assembly of wax and nanoparticles into hierarchical structures. Chem. Eng. J. 2020, 396, 125230. [Google Scholar] [CrossRef]
  40. Tong, J.; Liu, S.; Peng, R.; Sun, H.; Jiang, S. Development of a micro/nano composite super-hydrophobic silicon surface with nail-shaped texture/dual self-assembly monolayers and its wetting behavior. Appl. Surf. Sci. 2021, 544, 148803. [Google Scholar] [CrossRef]
  41. Etsion, I. State of the Art in Laser Surface Texturing. J. Tribol. 2005, 127, 248–253. [Google Scholar] [CrossRef]
  42. Qiu, T.Q.; Tien, C.L. Heat Transfer Mechanisms During Short-Pulse Laser Heating of Metals. J. Heat Transf. 1993, 115, 835–841. [Google Scholar] [CrossRef]
  43. De Zanet, A.; Casalegno, V.; Salvo, M. Laser surface texturing of ceramics and ceramic composite materials–A review. Ceram. Int. 2021, 47, 7307–7320. [Google Scholar] [CrossRef]
  44. Khan, S.A.; Boltaev, G.S.; Iqbal, M.; Kim, V.; Ganeev, R.A.; Alnaser, A.S. Ultrafast fiber laser-induced fabrication of superhydrophobic and self-cleaning metal surfaces. Appl. Surf. Sci. 2021, 542, 148560. [Google Scholar] [CrossRef]
  45. Dong, Z.; Sun, X.; Kong, D.; Chu, D.; Hu, Y.; Duan, J.-A. Spatial light modulated femtosecond laser ablated durable superhydrophobic copper mesh for oil-water separation and self-cleaning. Surf. Coat. Technol. 2020, 402, 126254. [Google Scholar] [CrossRef]
  46. Tian, G.; Zhang, Y.; Feng, X.; Hu, Y. Focus on Bioinspired Textured Surfaces toward Fluid Drag Reduction: Recent Progresses and Challenges. Adv. Eng. Mater. 2022, 24, 2100696. [Google Scholar] [CrossRef]
  47. Rong, W.; Zhang, H.; Zhang, T.; Mao, Z.; Liu, X.; Song, K. Drag Reduction Using Lubricant-Impregnated Anisotropic Slippery Surfaces Inspired by Bionic Fish Scale Surfaces Containing Micro-/Nanostructured Arrays. Adv. Eng. Mater. 2021, 23, 2000821. [Google Scholar] [CrossRef]
  48. Yang, X.; Zhuang, K.; Lu, Y.; Wang, X. Creation of Topological Ultraslippery Surfaces for Droplet Motion Control. ACS Nano 2020, 15, 2589–2599. [Google Scholar] [CrossRef]
  49. Yang, X.; Li, Y.; Zheng, H.; Lu, Y. Saturated Surface Charging on Micro/Nanoporous Polytetrafluoroethylene for Droplet Manipulation. ACS Appl. Nano Mater. 2022, 5, 3342–3351. [Google Scholar] [CrossRef]
  50. Zhu, M.; Liu, Y.; Chen, M.; Xu, Z.; Li, L.; Zhou, Y. Metal mesh-based special wettability materials for oil-water separation: A review of the recent development. J. Pet. Sci. Eng. 2021, 205, 108889. [Google Scholar] [CrossRef]
  51. Liu, Y.-Q.; Han, D.-D.; Jiao, Z.-Z.; Liu, Y.; Jiang, H.-B.; Wu, X.-H.; Ding, H.; Zhang, Y.-L.; Sun, H.-B. Laser-structured Janus wire mesh for efficient oil–water separation. Nanoscale 2017, 9, 17933–17938. [Google Scholar] [CrossRef] [PubMed]
  52. Volpe, A.; Gaudiuso, C.; Ancona, A. Laser Fabrication of Anti-Icing Surfaces: A Review. Materials 2020, 13, 5692. [Google Scholar] [CrossRef] [PubMed]
  53. Zheng, H.; Chang, S.; Ma, G.; Wang, S. Anti-icing performance of superhydrophobic surface fabricated by femtosecond laser composited dual-layers coating. Energy Build. 2020, 223, 110175. [Google Scholar] [CrossRef]
  54. Shen, X.; Yang, L.; Fan, S.; Yang, Q.; Wu, W.; Zhang, B. Colorful and superhydrophobic titanium surfaces textured by obliquely incident femtosecond laser induced micro/nano structures. Opt. Commun. 2020, 466, 125687. [Google Scholar] [CrossRef]
  55. Pan, R.; Zhang, H.; Zhong, M. Triple-Scale Superhydrophobic Surface with Excellent Anti-Icing and Icephobic Performance via Ultrafast Laser Hybrid Fabrication. ACS Appl. Mater. Interfaces 2020, 13, 1743–1753. [Google Scholar] [CrossRef]
  56. Hu, M.; Xu, K.; Zhang, Z.; Gao, J.; Wang, T.; Dai, X.; Zhu, H. Realizing of Anisotropic Wetting and Superhydrophobicity on Stainless Steel Surface by Combining Picosecond Laser Ablation and Electrodeposition. Steel Res. Int. 2021, 92, 2100219. [Google Scholar] [CrossRef]
  57. Boinovich, L.B.; Emelyanenko, K.A.; Domantovsky, A.G.; Chulkova, E.V.; Shiryaev, A.A.; Emelyanenko, A.M. Pulsed Laser Induced Triple Layer Copper Oxide Structure for Durable Polyfunctionality of Superhydrophobic Coatings. Adv. Mater. Interfaces 2018, 5, 1801099. [Google Scholar] [CrossRef]
  58. Yang, Z.; Liu, X.; Tian, Y. Hybrid Laser Ablation and Chemical Modification for Fast Fabrication of Bio-inspired Super-hydrophobic Surface with Excellent Self-cleaning, Stability and Corrosion Resistance. J. Bionic Eng. 2019, 16, 13–26. [Google Scholar] [CrossRef] [Green Version]
  59. Sanjay, S.L.; Annaso, B.G.; Chavan, S.M.; Rajiv, S.V. Recent progress in preparation of superhydrophobic surfaces: A review. J. Surf. Eng. Mater. Adv. Technol. 2012, 2012, 76–94. [Google Scholar]
  60. Zeng, Q.; Zhou, H.; Huang, J.; Guo, Z. Review on the recent development of durable superhydrophobic materials for practical applications. Nanoscale 2021, 13, 11734–11764. [Google Scholar] [CrossRef]
  61. Quan, Y.-Y.; Chen, Z.; Lai, Y.; Huang, Z.-S.; Li, H. Recent advances in fabricating durable superhydrophobic surfaces: A review in the aspects of structures and materials. Mater. Chem. Front. 2021, 5, 1655–1682. [Google Scholar] [CrossRef]
  62. Wang, S.; Liu, K.; Yao, X.; Jiang, L. Bioinspired Surfaces with Superwettability: New Insight on Theory, Design, and Applications. Chem. Rev. 2015, 115, 8230–8293. [Google Scholar] [CrossRef] [PubMed]
  63. Thomas, Y. An essay on the cohesion of fluids. Philos. Trans. R. Soc. Lond. 1805, 95, 65–87. [Google Scholar]
  64. Wenzel, R.N. Resistance of Solid Surfaces to Wetting by Water. Ind. Eng. Chem. 1936, 28, 988–994. [Google Scholar] [CrossRef]
  65. Cassie, A.B.D.; Baxter, S. Wettability of porous surfaces. Trans. Faraday Soc. 1944, 40, 546–551. [Google Scholar] [CrossRef]
  66. Erbil, H.Y.; Cansoy, C.E. Range of Applicability of the Wenzel and Cassie−Baxter Equations for Superhydrophobic Surfaces. Langmuir 2009, 25, 14135–14145. [Google Scholar] [CrossRef] [PubMed]
  67. Koishi, T.; Yasuoka, K.; Fujikawa, S.; Ebisuzaki, T.; Zeng, X.C. Coexistence and transition between Cassie and Wenzel state on pillared hydrophobic surface. Proc. Natl. Acad. Sci. USA 2009, 106, 8435–8440. [Google Scholar] [CrossRef] [Green Version]
  68. Vrancken, R.J.; Kusumaatmaja, H.; Hermans, K.; Prenen, A.M.; Pierre-Louis, O.; Bastiaansen, C.W.M.; Broer, D.J. Fully Reversible Transition from Wenzel to Cassie−Baxter States on Corrugated Superhydrophobic Surfaces. Langmuir 2010, 26, 3335–3341. [Google Scholar] [CrossRef]
  69. Arzt, E.; Quan, H.; McMeeking, R.M.; Hensel, R. Functional surface microstructures inspired by nature–From adhesion and wetting principles to sustainable new devices. Prog. Mater. Sci. 2021, 120, 100823. [Google Scholar] [CrossRef]
  70. Wang, D.; Huang, J.; Guo, Z. Tomato-lotus inspired edible superhydrophobic artificial lotus leaf. Chem. Eng. J. 2020, 400, 125883. [Google Scholar] [CrossRef]
  71. Yang, J.; Long, F.; Wang, R.; Zhang, X.; Yang, Y.; Hu, W.; Liu, L. Design of mechanical robust superhydrophobic Cu coatings with excellent corrosion resistance and self-cleaning performance inspired by lotus leaf. Colloids Surf. A Physicochem. Eng. Asp. 2021, 627, 127154. [Google Scholar] [CrossRef]
  72. Neinhuis, W.B. Purity of the sacred lotus, or escape from contamination in biological surfaces. Planta 1997, 202, 1–8. [Google Scholar]
  73. Yao, J.; JianNan, W.; YanHao, Y.; Han, Y.; Ying, X. Biomimetic fabrication and characterization of an artificial rice leaf surface with anisotropic wetting. Chin. Sci. Bull. 2012, 57, 2631–2634. [Google Scholar] [CrossRef]
  74. Yang, L.; Xiaodong, S.; Qin, Y.; Jiquan, L.; Daoyuan, L.; Jihone, D.; Bing, Z.; Shuqian, F. Fabrication of biomimetic anisotropic super-hydrophobic surface with rice leaf-like structures by femtosecond laser. Opt. Mater. 2020, 112, 110740. [Google Scholar] [CrossRef]
  75. Gao, F.; Yao, Y.; Wang, W.; Wang, X.; Li, L.; Zhuang, Q.; Lin, S. Light-Driven Transformation of Bio-Inspired Superhydrophobic Structure via Reconfigurable PAzoMA Microarrays: From Lotus Leaf to Rice Leaf. Macromolecules 2018, 51, 2742–2749. [Google Scholar] [CrossRef]
  76. Wang, X.; Dai, L.; Jiao, N.; Tung, S.; Liu, L. Superhydrophobic photothermal graphene composites and their functional applications in microrobots swimming at the air/water interface. Chem. Eng. J. 2021, 422, 129394. [Google Scholar] [CrossRef]
  77. Zhao, J.; Sun, R.; Liu, C.; Mo, J. Application of ZnO/epoxy resin superhydrophobic coating for buoyancy enhancement and drag reduction. Colloids Surf. A Physicochem. Eng. Asp. 2022, 651, 129714. [Google Scholar] [CrossRef]
  78. Wang, X.; Lin, D.; Zhou, Y.; Jiao, N.; Tung, S.; Liu, L. Multistimuli-Responsive Hydroplaning Superhydrophobic Microrobots with Programmable Motion and Multifunctional Applications. ACS Nano 2022, 16, 14895–14906. [Google Scholar] [CrossRef]
  79. Gao, X.; Jiang, L. Water-repellent legs of water striders. Nature 2004, 432, 36. [Google Scholar] [CrossRef]
  80. Shao, Y.; Zhao, J.; Fan, Y.; Wan, Z.; Lu, L.; Zhang, Z.; Ming, W.; Ren, L. Shape memory superhydrophobic surface with switchable transition between “Lotus Effect” to “Rose Petal Effect”. Chem. Eng. J. 2020, 382, 122989. [Google Scholar] [CrossRef]
  81. Li, Z.; Cao, M.; Li, P.; Zhao, Y.; Bai, H.; Wu, Y.; Jiang, L.J. Surface-embedding of functional micro-/nanoparticles for achieving versatile superhydrophobic interfaces. Matter 2019, 1, 661–673. [Google Scholar] [CrossRef] [Green Version]
  82. Zheng, Y.; Gao, X.; Jiang, L. Directional adhesion of superhydrophobic butterfly wings. Soft Matter 2006, 3, 178–182. [Google Scholar] [CrossRef] [PubMed]
  83. Ju, J.; Bai, H.; Zheng, Y.; Zhao, T.; Fang, R.; Jiang, L. A multi-structural and multi-functional integrated fog collection system in cactus. Nat. Commun. 2012, 3, 1–6. [Google Scholar] [CrossRef] [PubMed]
  84. Li, K.; Xie, Y.; Lei, J.; Zhang, S.; Liu, Z.; Lu, L. An inspiration from purple orchid leaves: Surface characteristics and wettability of nanoscale organometallic coatings electrodeposited on laser-patterned microstructures. Surf. Coat. Technol. 2021, 427, 127817. [Google Scholar] [CrossRef]
  85. Feng, L.; Li, S.; Li, Y.; Li, H.; Zhang, L.; Zhai, J.; Song, Y.; Liu, B.; Jiang, L.; Zhu, D. Super-Hydrophobic Surfaces: From Natural to Artificial. Adv. Mater. 2002, 14, 1857–1860. [Google Scholar] [CrossRef]
  86. Guo, Z.; Liu, W. Biomimic from the superhydrophobic plant leaves in nature: Binary structure and unitary structure. Plant Sci. 2007, 172, 1103–1112. [Google Scholar] [CrossRef]
  87. Feng, L.; Zhang, Y.; Xi, J.; Zhu, Y.; Wang, N.; Xia, A.F.; Jiang, L. Petal Effect: A Superhydrophobic State with High Adhesive Force. Langmuir 2008, 24, 4114–4119. [Google Scholar] [CrossRef] [PubMed]
  88. He, X.; Li, G.; Zhang, Y.; Lai, X.; Zhou, M.; Xiao, L.; Tang, X.; Hu, Y.; Liu, H.; Yang, Y.; et al. Bioinspired functional glass integrated with multiplex repellency ability from laser-patterned hexagonal texturing. Chem. Eng. J. 2021, 416, 129113. [Google Scholar] [CrossRef]
  89. Helbig, R.; Nickerl, J.; Neinhuis, C.; Werner, C. Smart Skin Patterns Protect Springtails. PLoS ONE 2011, 6, e25105. [Google Scholar] [CrossRef] [Green Version]
  90. Hensel, R.; Finn, A.; Helbig, R.; Braun, H.G.; Neinhuis, C.; Fischer, W.J.; Werner, C. Biologically Inspired Omniphobic Surfaces by Reverse Imprint Lithography. Adv. Mater. 2014, 26, 2029–2033. [Google Scholar] [CrossRef]
  91. Mingxia, S.; Watson, G.S.; Yongmei, Z.; Watson, J.A.; Liang, A. Wetting properties on nanostructured surfaces of cicada wings. J. Exp. Biol. 2009, 212, 3148–3155. [Google Scholar]
  92. Bellanger, H.; Darmanin, T.; de Givenchy, E.; Guittard, F. Chemical and Physical Pathways for the Preparation of Superoleophobic Surfaces and Related Wetting Theories. Chem. Rev. 2014, 114, 2694–2716. [Google Scholar] [CrossRef] [PubMed]
  93. Li, X.; Jiang, Y.; Jiang, Z.; Li, Y.; Wen, C.; Zhang, D.; Lian, J.; Zhang, Z. Improvement of corrosion resistance of H59 brass through fabricating superhydrophobic surface using laser ablation and heating treatment. Corros. Sci. 2020, 180, 109186. [Google Scholar] [CrossRef]
  94. Wang, M.; Liu, Q.; Zhang, H.; Wang, C.; Wang, L.; Xiang, B.; Fan, Y.; Guo, C.F.; Ruan, S. Laser Direct Writing of Tree-Shaped Hierarchical Cones on a Superhydrophobic Film for High-Efficiency Water Collection. ACS Appl. Mater. Interfaces 2017, 9, 29248–29254. [Google Scholar] [CrossRef]
  95. Ta, V.D.; Dunn, A.; Wasley, T.J.; Li, J.; Kay, R.W.; Stringer, J.; Smith, P.J.; Esenturk, E.; Connaughton, C.; Shephard, J.D. Laser textured superhydrophobic surfaces and their applications for homogeneous spot deposition. Appl. Surf. Sci. 2016, 365, 153–159. [Google Scholar] [CrossRef] [Green Version]
  96. Zhu, Z.; Wu, J.-R.; Wu, Z.-P.; Wu, T.-N.; He, Y.-C.; Yin, K. Femtosecond laser micro/nano fabrication for bioinspired superhydrophobic or underwater superoleophobic surfaces. J. Cent. South Univ. 2021, 28, 3882–3906. [Google Scholar] [CrossRef]
  97. Lian, Z.; Xu, J.; Yu, P.; Yu, Z.; Wang, Z.; Yu, H. Oil-Repellent and Corrosion Resistance Properties of Superhydrophobic and Superoleophobic Aluminum Alloy Surfaces Based on Nanosecond Laser-Textured Treatment. Met. Mater. Int. 2020, 26, 1603–1610. [Google Scholar] [CrossRef]
  98. Wang, T.; Zhu, H.; Zhang, Z.; Gao, J.; Wu, Y.; Hu, M.; Xu, K. Preparing of superamphiphobic surface by fabricating hierarchical nano re-entrant pyramids on micro-cones using a combined laser-electrochemistry method. Surf. Interfaces 2021, 24, 101112. [Google Scholar] [CrossRef]
  99. Long, J.; Wu, Y.; Gong, D.; Fan, P.; Jiang, D.; Zhang, H.; Zhong, M. Femtosecond laser fabricated superhydrophobic copper surfaces and their anti-icing properties. Chin. J. Lasers 2015, 42, 0706002. [Google Scholar] [CrossRef]
  100. Hauschwitz, P.; Jagdheesh, R.; Rostohar, D.; Brajer, J.; Kopeček, J.; Jiřícek, P.; Houdková, J.; Mocek, T. Hydrophilic to ultrahydrophobic transition of Al 7075 by affordable ns fiber laser and vacuum processing. Appl. Surf. Sci. 2020, 505, 144523. [Google Scholar] [CrossRef]
  101. Rajab, F.H.; Liu, Z.; Li, L. Long term superhydrophobic and hybrid superhydrophobic/superhydrophilic surfaces produced by laser surface micro/nano surface structuring. Appl. Surf. Sci. 2018, 466, 808–821. [Google Scholar] [CrossRef] [Green Version]
  102. Gaddam, A.; Sharma, H.; Karkantonis, T.; Dimov, S. Anti-icing properties of femtosecond laser-induced nano and multiscale topographies. Appl. Surf. Sci. 2021, 552, 149443. [Google Scholar] [CrossRef]
  103. Huerta-Murillo, D.; García-Girón, A.; Romano, J.-M.; Cardoso, J.T.; Cordovilla, F.; Walker, M.; Dimov, S.S.; Ocaña, J.L. Wettability modification of laser-fabricated hierarchical surface structures in Ti-6Al-4V titanium alloy. Appl. Surf. Sci. 2019, 463, 838–846. [Google Scholar] [CrossRef]
  104. Malinauskas, M.; Žukauskas, A.; Hasegawa, S.; Hayasaki, Y.; Mizeikis, V.; Buividas, R.; Juodkazis, S. Ultrafast laser processing of materials: From science to industry. Light. Sci. Appl. 2016, 5, e16133. [Google Scholar] [CrossRef] [PubMed]
  105. Phillips, K.C.; Gandhi, H.H.; Mazur, E.; Sundaram, S.K. Ultrafast laser processing of materials: A review. Adv. Opt. Photon- 2015, 7, 684–712. [Google Scholar] [CrossRef]
  106. Liu, B.; Wang, W.; Jiang, G.; Mei, X.; Wang, Z.; Wang, K.; Cui, J. Study on hierarchical structured PDMS for surface super-hydrophobicity using imprinting with ultrafast laser structured models. Appl. Surf. Sci. 2016, 364, 528–538. [Google Scholar] [CrossRef]
  107. Lee, Y.; Chung, Y.-W.; Park, J.; Park, K.; Seo, Y.; Hong, S.-N.; Lee, S.H.; Jeon, H.; Seo, J. Lubricant-infused directly engraved nano-microstructures for mechanically durable endoscope lens with anti-biofouling and anti-fogging properties. Sci. Rep. 2020, 10, 17454. [Google Scholar] [CrossRef]
  108. Karkantonis, T.; Gaddam, A.; Sharma, H.; Cummins, G.; See, T.L.; Dimov, S. Laser-Enabled Surface Treatment of Disposable Endoscope Lens with Superior Antifouling and Optical Properties. Langmuir 2022, 38, 11392–11405. [Google Scholar] [CrossRef]
  109. Fan, W.; Qian, J.; Bai, F.; Li, Y.; Wang, C.; Zhao, Q.-Z. A facile method to fabricate superamphiphobic polytetrafluoroethylene surface by femtosecond laser pulses. Chem. Phys. Lett. 2016, 644, 261–266. [Google Scholar] [CrossRef]
  110. Samanta, A.; Huang, W.; Chaudhry, H.; Wang, Q.; Shaw, S.K.; Ding, H. Design of Chemical Surface Treatment for Laser-Textured Metal Alloys to Achieve Extreme Wetting Behavior. ACS Appl. Mater. Interfaces 2020, 12, 18032–18045. [Google Scholar] [CrossRef]
  111. Ma, W.; Zhao, M.; Zhao, J.; Guo, Z.; Cui, Y.; Li, H. Preparation of superamphiphobic stainless steel surface by laser and perfluorooctanoic acid treatment. Opt. Laser Technol. 2022, 154, 108321. [Google Scholar] [CrossRef]
  112. Dong, C.; Gu, Y.; Zhong, M.; Li, L.; Sezer, K.; Ma, M.; Liu, W. Fabrication of superhydrophobic Cu surfaces with tunable regular micro and random nano-scale structures by hybrid laser texture and chemical etching. J. Mater. Process. Technol. 2011, 211, 1234–1240. [Google Scholar] [CrossRef]
  113. Liu, Y.; Liu, J.; Li, S.; Liu, J.; Han, Z.; Ren, L. Biomimetic Superhydrophobic Surface of High Adhesion Fabricated with Micronano Binary Structure on Aluminum Alloy. ACS Appl. Mater. Interfaces 2013, 5, 8907–8914. [Google Scholar] [CrossRef] [PubMed]
  114. Wang, M.; Yang, Z.; Yang, C.; Zhang, D.; Tian, Y.; Liu, X. The investigation of mechanical and thermal properties of super-hydrophobic nitinol surfaces fabricated by hybrid methods of laser irradiation and carbon ion implantation. Appl. Surf. Sci. 2020, 527, 146889. [Google Scholar] [CrossRef]
  115. Li, Y.; Li, M.; Utaka, Y.; Yang, C.; Wang, M. Effect of copper surface modification applied by combined modification of metal vapor vacuum arc ion implantation and laser texturing on anti-frosting property. Energy Build. 2020, 223, 110132. [Google Scholar] [CrossRef]
  116. Li, B.-J.; Huang, L.-J.; Ren, N.-F.; Kong, X. Laser ablation processing of zinc sheets in hydrogen peroxide solution for preparing hydrophobic microstructured surfaces. Mater. Lett. 2016, 164, 384–387. [Google Scholar] [CrossRef]
  117. Yan, X.; Huang, Z.; Sett, S.; Oh, J.; Cha, H.; Li, L.; Feng, L.; Wu, Y.; Zhao, C.; Orejon, D.; et al. Atmosphere-Mediated Superhydrophobicity of Rationally Designed Micro/Nanostructured Surfaces. ACS Nano 2019, 13, 4160–4173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Wang, D.; Wang, X.; Liu, X.; Zhou, F. Engineering a Titanium Surface with Controllable Oleophobicity and Switchable Oil Adhesion. J. Phys. Chem. C 2010, 114, 9938–9944. [Google Scholar] [CrossRef]
  119. Kwon, M.H.; Shin, H.S.; Chu, C.N. Fabrication of a super-hydrophobic surface on metal using laser ablation and electrodeposition. Appl. Surf. Sci. 2014, 288, 222–228. [Google Scholar] [CrossRef]
  120. Song, J.; Pan, W.; Wang, K.; Chen, F.; Sun, Y. Fabrication of micro-re-entrant structures by liquid/gas interface shape-regulated electrochemical deposition. Int. J. Mach. Tools Manuf. 2020, 159, 103637. [Google Scholar] [CrossRef]
  121. Hu, L.; Zhang, L.; Wang, D.; Lin, X.; Chen, Y. Fabrication of biomimetic superhydrophobic surface based on nanosecond laser-treated titanium alloy surface and organic polysilazane composite coating. Colloids Surf. A Physicochem. Eng. Asp. 2018, 555, 515–524. [Google Scholar] [CrossRef]
  122. Abele, L.; Jäger, A.; Schulz, W.; Ruck, S.; Riegel, H.; Sörgel, T.; Albrecht, J. Superoleophobic surfaces via functionalization of electrophoretic deposited SiO2 spheres on smart aluminum substrates. Appl. Surf. Sci. 2019, 490, 56–60. [Google Scholar] [CrossRef]
  123. Han, J.; Cai, M.; Lin, Y.; Liu, W.; Luo, X.; Zhang, H.; Zhong, M. 3D re-entrant nanograss on microcones for durable superamphiphobic surfaces via laser-chemical hybrid method. Appl. Surf. Sci. 2018, 456, 726–736. [Google Scholar] [CrossRef]
  124. Mousavi, S.; Pitchumani, R. A study of corrosion on electrodeposited superhydrophobic copper surfaces. Corros. Sci. 2021, 186, 109420. [Google Scholar] [CrossRef]
  125. Liu, J.; Fang, X.; Zhu, C.; Xing, X.; Cui, G.; Li, Z. Fabrication of superhydrophobic coatings for corrosion protection by electrodeposition: A comprehensive review. Colloids Surf. A: Physicochem. Eng. Asp. 2020, 607, 125498. [Google Scholar] [CrossRef]
  126. Yan, C.; Jiang, P.; Jia, X.; Wang, X. 3D printing of bioinspired textured surfaces with superamphiphobicity. Nanoscale 2020, 12, 2924–2938. [Google Scholar] [CrossRef]
  127. Zhang, H.; Gu, D.; Ma, C.; Xia, M.; Guo, M. Surface wettability and superhydrophobic characteristics of Ni-based nanocomposites fabricated by selective laser melting. Appl. Surf. Sci. 2019, 476, 151–160. [Google Scholar] [CrossRef]
  128. Huang, W.; Nelson, B.; Tian, S.; Ordikhani-Seyedlar, R.; Auyeung, R.C.; Samanta, A.; Hu, H.; Shaw, S.; Lamuta, C.; Ding, H. Superhydrophobic surface processing for metal 3D printed parts. Appl. Mater. Today 2022, 29, 101630. [Google Scholar] [CrossRef]
  129. Dong, Z.; Schumann, M.F.; Hokkanen, M.J.; Chang, B.; Welle, A.; Zhou, Q.; Ras, R.H.; Xu, Z.; Wegener, M.; Levkin, P.A. Superoleophobic slippery Lubricant-Infused surfaces: Combining two extremes in the same surface. Adv. Mater. 2018, 30, 1803890. [Google Scholar] [CrossRef] [Green Version]
  130. Lin, Z.-H.; Cheng, G.; Lee, S.; Pradel, K.C.; Wang, Z.L. Harvesting Water Drop Energy by a Sequential Contact-Electrification and Electrostatic-Induction Process. Adv. Mater. 2014, 26, 4690–4696. [Google Scholar] [CrossRef]
  131. Accardo, A.; Mecarini, F.; Leoncini, M.; Brandi, F.; Di Cola, E.; Burghammer, M.; Riekel, C.; Di Fabrizio, E. Fast, active droplet interaction: Coalescence and reactive mixing controlled by electrowetting on a superhydrophobic surface. Lab A Chip 2013, 13, 332–335. [Google Scholar] [CrossRef] [PubMed]
  132. Huang, X.; Mutlu, H.; Theato, P. A bioinspired hierarchical underwater superoleophobic surface with reversible pH response. Adv. Mater. Interfaces 2020, 7, 2000101. [Google Scholar] [CrossRef] [Green Version]
  133. Dong, S.; Wang, Z.; An, L.; Li, Y.; Wang, B.; Ji, H.; Wang, H. Facile fabrication of a superhydrophobic surface with robust micro-/nanoscale hierarchical structures on titanium substrate. Nanomaterials 2020, 10, 1509. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Wetting characterization: (a) Contact angle (θCA), (b) Rolling angle (θRA).
Figure 1. Wetting characterization: (a) Contact angle (θCA), (b) Rolling angle (θRA).
Crystals 13 00020 g001
Figure 2. Four wetting conditions.
Figure 2. Four wetting conditions.
Crystals 13 00020 g002
Figure 3. Schematic diagram of laser-prepared microstructures.
Figure 3. Schematic diagram of laser-prepared microstructures.
Crystals 13 00020 g003
Figure 4. Industrial applications of functional surfaces prepared by laser micromachining.
Figure 4. Industrial applications of functional surfaces prepared by laser micromachining.
Crystals 13 00020 g004
Figure 5. Schematics of wetting models: (a) Young model, (b) Wenzel model, (c) Cassie–Baxter model.
Figure 5. Schematics of wetting models: (a) Young model, (b) Wenzel model, (c) Cassie–Baxter model.
Crystals 13 00020 g005
Figure 6. Diagram of organisms with wetting properties and microstructures: (a) Lotus leaf, (b) SEM of lotus leaf [86], (c) Rice leaf, (d) SEM of rice leaf [86], (e) Rose petal, (f) SEM of Rose petal [87], (g) Water strider, (h) SEM of Water strider, (i) Butterfly wing, (j) SEM of Butterfly wing, (k) Mosquito compound eye, (l) Mosquito compound eye [88].
Figure 6. Diagram of organisms with wetting properties and microstructures: (a) Lotus leaf, (b) SEM of lotus leaf [86], (c) Rice leaf, (d) SEM of rice leaf [86], (e) Rose petal, (f) SEM of Rose petal [87], (g) Water strider, (h) SEM of Water strider, (i) Butterfly wing, (j) SEM of Butterfly wing, (k) Mosquito compound eye, (l) Mosquito compound eye [88].
Crystals 13 00020 g006
Figure 7. Organisms with superamphiphobic wetting properties: (a) SEM images of Springtail [89], (b) Cicada [91], (c) Surface microstructure of different cicadas [91].
Figure 7. Organisms with superamphiphobic wetting properties: (a) SEM images of Springtail [89], (b) Cicada [91], (c) Surface microstructure of different cicadas [91].
Crystals 13 00020 g007
Figure 8. Laser-fabricated superhydrophobic surfaces: (a) Preparation of superhydrophobic surfaces by laser-chemical modification [99], (b) Laser and vacuum treatments achieve superhydrophobic surfaces [100], (c) Superhydrophobic or superoleophobic surfaces prepared by the nanosecond and picosecond laser [101].
Figure 8. Laser-fabricated superhydrophobic surfaces: (a) Preparation of superhydrophobic surfaces by laser-chemical modification [99], (b) Laser and vacuum treatments achieve superhydrophobic surfaces [100], (c) Superhydrophobic or superoleophobic surfaces prepared by the nanosecond and picosecond laser [101].
Crystals 13 00020 g008
Figure 9. Process of preparing non-metallic superhydrophobic surface: (a) superhydrophobic surfaces of PDMS [106], (b) Process of preparing superhydrophobic surfaces for lenses [107].
Figure 9. Process of preparing non-metallic superhydrophobic surface: (a) superhydrophobic surfaces of PDMS [106], (b) Process of preparing superhydrophobic surfaces for lenses [107].
Crystals 13 00020 g009
Figure 10. Laser preparation of superamphiphobic surface surfaces: (a) PTFE microgrooves machined by femtosecond laser [109], (b) Aluminum alloy superamphiphobic surface prepared by laser etching [91], (c) Superamphiphobic aluminum alloy surface in FOTS, FDTS and FDDTS solutions [110], (d) Laser etching of reticulated stainless-steel surface [111].
Figure 10. Laser preparation of superamphiphobic surface surfaces: (a) PTFE microgrooves machined by femtosecond laser [109], (b) Aluminum alloy superamphiphobic surface prepared by laser etching [91], (c) Superamphiphobic aluminum alloy surface in FOTS, FDTS and FDDTS solutions [110], (d) Laser etching of reticulated stainless-steel surface [111].
Crystals 13 00020 g010
Figure 11. Preparation of superamphiphobic surface by laser and its hybrid methods: (a) Laser texturing of aluminum alloy surface and deposition of SiO2 micro-spheres [122], (b) Laser-chemical bath deposition [123], (c) Laser-electrochemical hybrid fabrication [98], (d) Laser-electrochemical hybrid fabrication of roof structures [120].
Figure 11. Preparation of superamphiphobic surface by laser and its hybrid methods: (a) Laser texturing of aluminum alloy surface and deposition of SiO2 micro-spheres [122], (b) Laser-chemical bath deposition [123], (c) Laser-electrochemical hybrid fabrication [98], (d) Laser-electrochemical hybrid fabrication of roof structures [120].
Crystals 13 00020 g011
Table 1. Laser and its hybrid processing for the preparation of superhydrophobic and superamphiphobic surfaces.
Table 1. Laser and its hybrid processing for the preparation of superhydrophobic and superamphiphobic surfaces.
Method 1Method 2Surface ModificationMaterialsStructural FeaturesWetting PropertiesRef.
fs laser\fluorinatedCuperiodic micron-sized bumps and depressionssuperhydrophobicity[99]
ns laser\VacuumAl alloyarray of micro-conesSuperhydrophobicity,
θwCA = 170°, θwRA = 5°
[100]
ps laser
ns laser
\fluorinated316ssperiodic structuresuperhydrophobicity[101]
fs laser\lubricant430ssperiodic structuresuperhydrophobicity,
θwCA = 162°
[102]
ns laserfs laserpolyethyleneTi-6Al-4Vmicro-pillars, periodic structuresuperhydrophobicity,
θwCA > 160°
[103]
ps laser\\PDMSperiodic groovesuperhydrophobicity,
θwCA = 170°
[106]
fs laser\fluorinatedbk7
glass
periodic wavesuperhydrophobicity[107]
fs laserhot embossinglubricant316ss, polycarbonate,
cyclic olefin
periodic structuresuperhydrophobicity[108]
fs laser\\PTFEmicrogrooves, layered submicron structuressuperamphiphobicity,
θWCA = 158.9°,
θOCA = 153.4°
[109]
ns laser\Heat treatment,
fluorinated
Al alloynet microtexturesuperamphiphobicity,
θWCA = 155.3°,
θOCA = 153.8°
[97]
ns laser\fluorinatedAl alloyperiodic groovesuperamphiphobicity[110]
ns laser\fluorinatedstainless steelnet microtexturesuperamphiphobicity,
θWCA = 165°,
θOCA = 155°
[111]
ns laseroxidationfluorinatedTiTiO2 nanotube arrayssuperamphiphobicity,
θWCA > 160°,
θOCA > 160°
[118]
ns laserelectrophoretic depositionfluorinatedAllayered structures, pits and micro-spheressuperamphiphobicity,
θOCA = 157°
[122]
fs laserchemical bath processingfluorinatedAl alloyMicro-cone surface,
nano-grass CuO
superamphiphobicity,
θWCA > 150°,
θOCA > 150°
[123]
ps laserelectrodepositionfluorinatedCucopper micro-cone, nickel conesuperamphiphobicity,
θWCA = 161°,
θOCA = 151°
[98]
ns laserelectrodepositionfluorinatedCuarray roof structuresuperamphiphobicity,
θWCA = 161°,
θOCA = 151°
[120]
laserALDfluorinatedphotoresistarray of T-shaped micro-pillarssuperamphiphobicity,
θWCA = 162°,
θoCA = 160°
[129]
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

Liu, Y.; Wu, M.; Guo, C.; Zhou, D.; Wu, Y.; Wu, Z.; Lu, H.; Zhang, H.; Zhang, Z. A Review on Preparation of Superhydrophobic and Superoleophobic Surface by Laser Micromachining and Its Hybrid Methods. Crystals 2023, 13, 20. https://doi.org/10.3390/cryst13010020

AMA Style

Liu Y, Wu M, Guo C, Zhou D, Wu Y, Wu Z, Lu H, Zhang H, Zhang Z. A Review on Preparation of Superhydrophobic and Superoleophobic Surface by Laser Micromachining and Its Hybrid Methods. Crystals. 2023; 13(1):20. https://doi.org/10.3390/cryst13010020

Chicago/Turabian Style

Liu, Yang, Mingyi Wu, Chunfang Guo, Dong Zhou, Yucheng Wu, Zhaozhi Wu, Haifei Lu, Hongmei Zhang, and Zhaoyang Zhang. 2023. "A Review on Preparation of Superhydrophobic and Superoleophobic Surface by Laser Micromachining and Its Hybrid Methods" Crystals 13, no. 1: 20. https://doi.org/10.3390/cryst13010020

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