Next Article in Journal
Preparation of Ti-Doped ZnO/Bi2O3 Nanofilm Heterojunction and Analysis of Microstructure and Photoelectric Properties
Next Article in Special Issue
The Atomistic Understanding of the Ice Recrystallization Inhibition Activity of Antifreeze Glycoproteins
Previous Article in Journal
Reactive Spark Plasma Sintering and Thermoelectric Properties of Zintl Semiconducting Ca14Si19 Compound
Previous Article in Special Issue
Computational Analysis of Hydrogen Bond Vibrations of Ice III in the Far-Infrared Band
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ordered/Disordered Structures of Water at Solid/Liquid Interfaces

1
School of Physical Science and Intelligent Engineering, Jining University, Qufu 273155, China
2
Division of Interfacial Water and Key Laboratory of Interfacial Physics and Technology, Shanghai Institute of Applied Physics, Chinese Academy of Sciences, Shanghai 201800, China
3
University of Chinese Academy of Sciences, Beijing 100049, China
4
College of Sciences, Shanghai University, Shanghai 200444, China
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(2), 263; https://doi.org/10.3390/cryst13020263
Submission received: 3 January 2023 / Revised: 26 January 2023 / Accepted: 30 January 2023 / Published: 2 February 2023
(This article belongs to the Special Issue Novel Ice Crystals)

Abstract

:
Experiments and theory have revealed versatile possible phases for adsorbed and confined water on two-dimensional solid surfaces, which are closely related to the aspects of various phenomena in physics, chemistry, biology, and tribology. In this review, we summarize our recent works showing that the different water phases with disordered and ordered structures can greatly affect surface wetting behavior, dielectric properties, and frictions. This includes the ordered phase of water structure that induces an unexpected phenomenon, an “ordered water monolayer that does not completely wet water”, at T = 300 K on the model’s surface and some real, solid material, together with the anomalous low dielectric properties due to ordered water.

1. Introduction

An understanding of the interfacial water phase [1,2,3,4,5,6,7,8,9] is necessary for the understanding of various physical, chemical, and biological processes, such as the hydration of solutes in water [10,11,12,13], water adsorption behavior on solid surfaces [14,15,16,17], fluids and their transportation across channels [18,19,20], electrochemical reactions [21,22], protein stability and folding [23,24,25,26,27], and molecular peptide self-assembly [28,29]. Experiments and theories have suggested versatile possible phases for adsorbed and confined water [30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48], which are closely related to the aspects of various phenomena in the material sciences, such as geology, biology, tribology, and nanotechnology [10,49,50,51]. Unlike bulk liquids, water molecules strongly confined to or adsorbed on a solid surface usually exhibit a particular molecular structure and dynamic behavior. Due to the new high-resolution experimental technique developed in recent decades, the ordered phase of water molecules with a particular structure has been revealed [51,52,53,54,55]. In 1995, Hu et al. [54] firstly experimentally discovered the ordered water layer absorbed in the vicinity of the mica surface at room temperature. Subsequent work [56] by simulations and sum-frequency spectroscopy experiments performed by Miranda et al. [57] confirmed this two-dimensional hydrogen-bond network. Recently, atomic resolution-ordered water molecules have been directly observed on various materials’ surface, thanks to high-resolution experimental techniques. Tetramer-ordered water molecules have been observed on the surface of NaCl(001) by Guo et al. [58] through the scanning tunneling microscope (STM) technique, at only 5 K. In 2019, Ma et al. found novel edge structures of 2D bilayer hexagonal ice appearing on the surface of Au(111) for the first time using noncontact atomic-force microscopy. Nanoconfinement, even in a hydrophobic space, can also induce ordered structures. Using high-resolution electron microscopy imaging, Geim et al. [59] experimentally observed the forms of square ice in between two graphene sheets at room temperature, a phase different from the conventional tetrahedral water structures forming about four hydrogen bonds (H-bonds). In theory, several novel phase transitions of confined water were predicted. In 1997, Koga et al. [60] predicted a new bilayer ice phase form with interlocking H-bonds at extremely high pressures, and a new phase transition occurred when water molecules were confined between certain nanoscale spaces. In 2009, a new phase transition [61] between bilayer liquid and trilayer heterogeneous water confined in two parallel, atomically hydrophobic walls was observed, as water density increased. Interestingly, Han et al. [62] observed that the first-order transition line of water molecules confined between hydrophobic space can connect with a continuous transition line, instead of terminating at a critical point. Kapil et al. [63] combined state-of-the-art computational approaches to perform a first-principles-level study on a water monolayer confined within a graphene-like channel. They observed that this water layer exhibited interestingly abundant and various phase behaviors, which was greatly dependent on the pressure and temperature. These results suggest that confined water exhibits versatile and complex phase diagrams. In respect of the biology, interfacial phase water that is strongly associated with biomolecules has been observed near the biological molecules [64,65,66,67,68], which may play key roles in the function and diffusion of the biomolecules.
In the present review, we summarize the recent works showing that the phase transition of water from disordered to ordered structures in the first layer to contact with surface can greatly affect surface wetting behavior, surface frictions, and dielectric properties. We highlight the distinct phases of water structures’ contact with the surfaces with different H-bond networks, which are mainly attributed to the matching or mismatching between the water structure and the atomic arrangement of a solid surface. The article is organized in four sections. In Section 2, we show the prediction that the hydrophobicity of an ordered water monolayer can be unexpectedly enhanced by ordered water structure at room temperature, which is referred to as an “ordered water monolayer that does not completely wet water”. In Section 3, we show the novel wetting behavior of water on the COOH matrix and SAM-(OH)2. In Section 4, we show the dependence of the ordered water structure on parallel dielectric permittivity. In the last section, we provide a summary and the future directions of the research.

2. Phase Transition from Disordered to Ordered Water Structures Induces an “Ordered Water Monolayer That Does Not Completely Wet Water” at Room Temperature on Solid Surfaces

The hydrophobic-like ice monolayer at cryogenic temperatures [69] was not expected at room temperature, while room temperature, strong, thermal fluctuations usually disturb H-bond networks in water. It is not difficult to rationalize that water at room temperature plays a greater role in daily life and applications, including surface friction on solid/water interfaces, water warming or cooling, solvation of solvent molecules, biological activities and functions, and oil mining underground. Since hydrogen bonds form among water molecules, water molecules are naturally and always completely wetting other water. Figure 1a presents our molecular dynamics (MD) prediction that a water droplet with a pronounced contact angle stands on a very thin (0.4 nm) water monolayer in the vicinity of an ionic solid surface at room temperature. This phenomenon has been referred to as an “ordered water monolayer that does not completely wet water” at room temperature [70,71,72,73,74,75]. In Figure 1b, we present a theoretical ionic model surface with a planar hexagonal charge pattern structure with charge values q at certain locations, while the surface is neutral in total.
This phenomenon is mainly attributed to the ordered phase water structures resulting from the high charge and special arrangement of the solid surface, which decrease the number of H-bonds in the inter-layer and increase the number of H-bonds in the intra-layer. Our calculation shows that the thermal conductivity of the ordered water monolayer would most likely resemble ice rather than liquid water, from a thermodynamics viewpoint [76]. The solid lattice structure, such as the surface bond length l, is fundamental to the phase transition of disordered to ordered water structures [70,74], which can transform the nonwetted to a completely wetted water layer. l increases from 0.142 to 0.16 nm or decreases from 0.142 to 0.12 nm; we observed there were no ordered water structures, whereas charge q was high enough (1.0 e). Our recent study showed that the ordered–disordered phase transition of interfacial water was considerably relevant to the charge dipole moment, production of both charge values, and the dipole length of the solid surface. In addition, the surface point defects [74], temperature [77], and curvature [78] can also induce the wetting transition from nonwetted to completely wetted due to the disruption of ordered water structures. In 2015, we observed a 25% friction reduction on the super-hydrophilic materials due to the ordered water molecular structures [79]. This further demonstrates the important role of ordered water structures in hydrodynamics. Recently, the water-induced friction reduction was experimentally verified by Ma et al. on TiO2 silica surfaces [80] and other solid surfaces [81]. This super-hydrophilic but low-friction surface may have great application potential in self-clearing materials [82,83,84,85], biomedical materials [86,87], and nano-coating materials of ships [82].
Recently, we applied this theorical framework to solve the long-debated question of understanding wetting behaviors on TiO2 surfaces [88]. Since the first experimental observation of the wetting transformation on a solid surface under UV exposure in 1997, the wetting of a TiO2 surface has been debated for decades [89,90]. We noted that the macroscopic contact angle was around 15–30° initially on a freshly prepared TiO2 surface, but increased to 60–70° under ambient conditions in the dark [83,89,90,91,92,93,94]. However, water droplets diffuse all over the film following UV exposure, leading to super-hydrophilic behavior with a contact angle of 0°. Then, several works determined that the existence of an intrinsically hydrophobic (oleophilic) region [89,90] or adsorbed hydrophobic hydrocarbon contamination [95,96] contributed to nonwetting behavior. In 2018, Diebold et al. claimed that the amphiphilic carboxylic acid monolayer on rutile TiO2(110)’s surface could induce hydrophobic behavior through the experimental method combining atomic-scale microscopy (AFM), a scanning tunneling microscope (STM), and X-ray photoelectron spectroscopy (XPS) [97]. In fact, molecular-ordered water structures were observed in some theoretical works by MD simulations [91,98] and experiments under vacuum conditions or at cryogenic temperatures [99,100,101,102,103,104,105]. However, the relationship between water’s structure and wetting behavior was not carefully considered. Intuitively, the interactions (>1.0 eV) between TiO2 and water [106] are strong; thus, in principle, a super-hydrophilic surface similar to a mica surface [54,56] is expected. We then speculated that the unexpected wetting with a large contact angle can be exactly solved using our previous model shown in Figure 1. We used MD simulations with a classical force field and neural network potential (NN-MD) to firstly identify a water droplet on an ordered water bilayer structure in the vicinity of a rutile TiO2 surface under ambient conditions (see Figure 2a,b). The ordered water structure decreased the number of H-bonds between the bilayer and water droplet, and thus created an obvious contact angle. UV exposure with a reduction in the quantity of adsorbed molecular water and an enhancement in the quantity of adsorbed dissociated water with −OH at the surface was evident. Therefore, we investigated the effect of 5% and 10% covering ratios of −OH groups on the wetting behavior of rutile TiO2(110)’s surface, with the contact angle of the water droplet on the water bilayer decreasing from 19° (5%) to 0° (10%). This can be ascribed to the disruption of the water bilayer H-bonds network, which transforms the hydrophobicity of the water bilayer to super-hydrophilic.
Since 2009, many research groups have provided proof of this similar phenomenon [107], suggesting it extensively exists on several real materials’ surfaces. The majority of these works utilized the physical mechanism of the phase transition of water from disordered to ordered structures; we proposed to explain their observations in theoretical simulations or experiments. In 2015, we observed that this phenomenon may exist on Pd(100) (with a contact angle value of 57°), Pt(100) (with a contact angle value of 53°), and Al(100) (with a contact angle value of 32°) surfaces, while it cannot be observed on other solid-metal surfaces, such as (110) and (111) surfaces (see Figure 3c,e,f) and the (100) surface of Ni with a large lattice constant [108]. We observed that the clear, ordered water structure (see Figure 3g) induced the wetting transition from a completely wetted to a nonwetted water monolayer. As shown in Figure 3h, for various surfaces with different metal surfaces with different lattice constants, four clear orientation preferences of the water dipole are observed for Pd(100) and Pt(100) surfaces at φ = 0°, 90°, 180°, and φ = 270°, respectively. This indicates that the formation of ordered water structures is induced by the matching between the lattice structure for Pd(100) and Pt(100) surfaces and water H-bonds. Interestingly, these ordered water structures show rhombic structures, different from the previous hexagonal structures [70].
In 2011, Rotenberg et al. [109] showed the wetting behavior of a talc surface dependent on humidity using MD simulations, i.e., hydrophilic at low humidity levels, while hydrophobic at high humidity levels. This large contact angle is consistent with experiments that show that the macroscopic contact angle is about 80°–85° [110]. Then, Phan et al. [111] observed the phenomenon of an “ordered water monolayer that does not completely wet water” on hydroxylated SiO2 (111) and Al2O3 surfaces utilizing MD simulations. They directly used the physical mechanism of the disordered–ordered phase transition of interfacial water structures to explain the observations. In 2021, we observed the wetting phenomenon with a droplet on composite structures formed by embedded water in the (111) surface of β-cristobalite hydroxylated silica [112]. In 2013, Limmer et al. [21] also observed the novel wetting phenomenon that an “ordered water monolayer that does not completely wet water” on the Pt(100) surface. Note that each water sample in the ordered monolayer tends to form about four H-bonds, which is different from the hexagonal structure of the ordered water monolayer with three H-bonds on our model surface [70,71,113].
In addition to the simulations, in respect of the experiments, Lützenkirchen et al. [114,115] observed an ice-like configuration of water molecules on the sapphire c-plane surface by using the sum frequency generation (SFG) technique, which was distinct from the bulk liquid water phase. It should be noted that the contact angles of the liquid water droplets on the solid surface were quite large [114]. In 2014, Lee et al. [98] used the mechanism inspired by our study to explain experimental superwetting under light-illumination conditions on Titania surfaces. In 2021, direct experimental evidence for the observation of this ordered water layer on surfaces, particularly on biomolecule and polymer surfaces, was validated at room temperature for a hydrophobic fluorinated polymer, such as polytetrafluoroethylene’s (PTFE’s) surface [116], by employing SFG vibrational spectroscopy. These experimental evidence provides the platform for future applications in terms of the related materials.

3. Ordered Phase of Composite Structures of Water Molecules Embedded into the Carboxylic Acid-Terminated Self-Assembled Monolayers (COOH-SAMs) and Hydroxyl-Terminated Self-Monolayer ((OH)2-SAMs) at Room Temperature

Carboxylic acid-terminated self-assembled monolayers have attracted considerable interest due to their wide applications in nanoscience and nanotechnology [117,118,119,120,121,122,123,124,125]. However, there were inconsistent values of water contact angles on COOH-SAMs, even after 25 years of study, which still puzzles researches regarding the surface water adsorption behavior on COOH-SAMs. We collected these contact angle values and observed that the values fell in the range of 0° to 50° from forty literature experiments [118,126,127,128,129]. Particularly, in 2011, James et al. [130] observed water droplets coexisting with a continuous few-angstrom-thin water layer on COOH-SAM using X-ray, neutron reflectometry, and atomic force microscopy (AFM) methods. Moreover, the adsorbed water molecules at the COOH-SAM surface showed novel adsorption behaviors, different from the conventional viewpoint. A similar situation can be observed for SAM-OH surfaces, which are usually regarded as super-hydrophilic.
We observed that water molecules can be embedded in the COOH matrix on COOH-SAMs with appropriate packing densities to form an ordered phase structure of an embedded water–COOH composite [131], which enhances the hydrophobicity of embedded water–COOH composite structures (see Figure 4a). Remarkably, a liquid water droplet with an apparent contact angle of approximately 34° stands on the embedded water–COOH composite [Figure 4a]. This phenomenon is caused by the water embedded in COOH groups forming an H-bond network, which leads to reduced H-bond numbers between the surface and water molecules above the composite structure. This can explain the experimental work by James et al. well. Figure 4b shows the contact angles of water droplets on COOH-SAMs as a function of density Σ. There is an angle plateau of ~35° in the mid-range and a value ~0° in the dense range. In the sparse range, we observed the contact angle θ initially decreased and then slightly increased as Σ increased. In 2013, Wang et al. [132] experimentally observed the phenomenon of water droplets coexisting with a nanoscale water layer formed on bovine serum albumin (BSA) when the IB was sealed at a low RH (15–25%) at room temperature for 3–5 days.
In addition to the hydrophilic group COOH, there are very few reports on how to obtain hydrophobicity on surfaces constituted by only utilizing hydrophilic groups/molecules. Our recent study determined that a typical hydrophilic -OH group can also induce a hydrophobic surface with large contact angles. In our study. we used five-carbon long alkyl chains, which were grafted on one end with two -OH groups exposed to water. The packing density (Σ) varied from 2.0 to 6.5 nm−2. Then, we performed MD simulations [133] to analyze the water droplet wetting behavior and calculated the contact angles of water droplets on (OH)2-SAMs. Figure 5a presents a representative snapshot of a water droplet on (OH)2-SAM for Σ = 4.5 nm−2, where the contact angle of the water droplet is 82°, indicating the hydrophobicity of (OH)2-SAM. However, the average number of H-bonds in between the neighboring OH groups is approximately 8.8 nm−2, suggesting that even the typical hydrophilic OH group can exhibit hydrophobic behavior. We also show that the contact angles of water droplets on (OH)2-SAMs depend on the function of packing density Σ, similar to COOH-SAM. Interestingly, the water molecules can be embedded into the OH groups to form composite structures with looser packing densities Σ. This composite structures also enhanced the hydrophobicity of (OH)2-SAM.

4. Effect of Ordered Phase of Water on the Dielectric Permittivity

Compared with other fluids, bulk water has a large, static, dielectric constant, which is very important in various aqueous environments, such as energy systems’ [134,135] biomolecule function [136,137] and ion-ion interactions [55,138,139]. In confinement or close to the interface, the dielectric properties of water become anisotropic [136,140,141,142,143,144,145,146]. For many decades, water’s dielectric permittivity has been intensively studied [136,140,141,142,143,144,145,147,148,149,150,151,152,153]. Recently, water’s dielectric permittivity in a perpendicular direction between hBN and graphene nanochannels has been measured [141]. For the lower values of interslab separation, an anomalously low perpendicular dielectric constant (as low as 2) was reported. In addition to nanoconfinement, the number density of water molecules [145,146] or surface wettability [140,154] were also reported to impact the dielectric permittivity of water close to solid surfaces.
Recent works have shown that an ordered water structure can form in the vicinity of some solid surfaces at room temperature, such as ionic model surfaces, metal surfaces [21,78,155,156], metal oxides [117,157], and clay surfaces [109]. Our MD simulations revealed that the parallel dielectric permittivity of interfacial water depends on solid-water interactions together with the interfacial water structure on various sold crystal faces [158]. In particular, the in-plane dielectric permittivity of ordered water structures on solid surfaces can be significantly reduced.
The different trends for the parallel dielectric constant of water ( ε Interfacial ) close to (100) and (111) surfaces versus the different surface-water interactions (f) are shown in Figure 6. For the fcc (100) surface, ε Interfacial initially increases from 60 to 102 and then reduces to 40. This sudden decrease in ε Interfacial originates from the ordered H-bond network (see Figure 3g), leading to the low amplitude of the dipolar fluctuation. This nonmonotonic behavior of ε Interfacial with f is quite different from previous reports that show that the more hydrophilicity on the surface, the larger the dielectric constant [145,146]. However, for the fcc (111) surface, ε Interfacial increases from 78 to 129 with f, in accordance with the previous reports [145,146]. Our work indicates that the crystal orientation of hydrophilic surfaces can significantly affect the parallel dielectric permittivity of interfacial water.

5. Summary

In the current review, we summarized the recent advances in how the phase transition of water from disordered to ordered structures in the first layer to contact with the solid surface affects a variety of surface properties, such as wetting behaviors, surface dielectric properties, and surface frictions. The ordered phases of water structure induce unexpected phenomenon, an “ordered water monolayer that does not completely wet water” at room temperature on ionic model solid surfaces, and metals, metal oxides, and minerals. We used this theorical framework to solve the long-debated question of understanding wetting behaviors on TiO2 surfaces. The ordered phase of composite structures of water molecules embedded in the carboxylic acid-terminated self-assembled monolayers (COOH-SAMs) enhances hydrophobicity at room temperature. Similarly, the hydrogen-bond network of hydroxyl-terminated self-assembled monolayers ((OH)2-SAMs) and water molecules embedded in (OH)2-SAMs also induce unexpected hydrophobicity at room temperatures. Particularly, those wetting behaviors previously described exhibit clear water droplets at a macroscopic level indicating hydrophobicity, but the water layer at a molecular level indicates hydrophilicity, which is termed “molecular-scale hydrophilicity” [159]. This transition from a disordered to ordered water structure is mainly attributed to the mismatching or matching between solid surface structures and water molecules’ surfaces. Our works show that the various microscopic water structures with variable quantity allocations of H-bond numbers between water molecules can regulate physical interactions, even macroscopic properties, i.e., wetting, friction reduction, and dielectric properties.
Despite the considerable efforts made, there are still many unanswered questions in this research field. Recent simulations and experimental works [88,98,114,115,116,130,132] have determined the phenomenon that an “ordered water monolayer that does not completely wet water”; however, whether there will be more materials found in this direction is still unknown. We noted that more direct results for ordered water structures that can lead to wetting behavior are still lacking at room temperature, which may be attributed to the lack of accuracy of the resolution in techniques at present. Subsequently, how these ordered water structures on solid metal or metal oxide surfaces affect surface properties, such as the catalysis, non-fouling, electric potential, and gas formation, is still in urgent need of research. For example, whether the ordered water hydrogen-bond network affects water transport through covalent organic framework channels [18,160], the evaporation of interfacial water [161] and lubrication strategy of hydrogels [162] is still an open question in terms of the applications. The so-called physical electric double layer (EDL) was usually induced by the charged surfaces when an aqueous electrolyte solution was present. However, the classical theory neglecting the interfacial molecular water structures is still unable to fully describe the observed phenomenon under the assumptions, such as no ion–ion correlations and the homogeneous dielectric continuum of water [163]. It should be noted that the quantum nuclei effect should be given more attention when introducing important effects on the hydrogen bonds and reorientation of water [164,165] spectra in terms of both the energy and line shapes [166].

Author Contributions

Conceptualization, C.Q. and C.W.; Methodology, C.W.; Software, C.Q. and C.L.; Validation, C.Q. and C.L.; formal analysis, C.Q.; investigation, C.Q. and C.W.; resources, C.W.; data curation, C.Q. and C.W.; writing—original draft preparation, C.Q. and C.W.; writing—review and editing, C.Q. and C.W.; supervision, C.W.; funding acquisition, C.Q. and C.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Science Foundation of China (Grant Nos. 12022508, 12074394, 11674345), the Key Research Program of the Chinese Academy of Sciences (QYZDJ-SSW-SLH019) and the Natural Science Foundation of Shandong Province of China (Grant No. ZR2022QA089).

Data Availability Statement

The data presented in this study are available on request.

Acknowledgments

Computations were performed on the Shanghai Supercomputer Center of China.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Stirnemann, G.; Castrillon, S.R.-V.; Hynes, J.T.; Rossky, P.J.; Debenedetti, P.G.; Laage, D. Non-monotonic dependence of water reorientation dynamics on surface hydrophilicity: Competing effects of the hydration structure and hydrogen-bond strength. Phys. Chem. Chem. Phys. 2011, 13, 19911–19917. [Google Scholar] [CrossRef] [PubMed]
  2. Giovambattista, N.; Debenedetti, P.G.; Rossky, P.J. Enhanced surface hydrophobicity by coupling of surface polarity and topography. Proc. Natl. Acad. Sci. USA 2009, 106, 15181–15185. [Google Scholar] [CrossRef] [PubMed]
  3. Godawat, R.; Jamadagni, S.N.; Garde, S. Characterizing hydrophobicity of interfaces by using cavity formation, solute binding, and water correlations. Proc. Natl. Acad. Sci. USA 2009, 106, 15119–15124. [Google Scholar] [CrossRef] [PubMed]
  4. Shi, G.; Liu, J.; Wang, C.; Song, B.; Tu, Y.; Hu, J.; Fang, H. Ion Enrichment on the Hydrophobic Carbon-based Surface in Aqueous Salt Solutions due to Cation-π Interactions. Sci. Rep. 2013, 3, 3436. [Google Scholar] [CrossRef]
  5. Qiu, Y.; Liu, Y.; Tu, Y.; Wang, C.; Xu, Y. Defect-Induced Wetting Behavior on Solid Polar Surfaces with Small Charge Dipole Length. J. Phys. Chem. C 2017, 121, 17365–17370. [Google Scholar] [CrossRef]
  6. Rego, N.B.; Patel, A.J. Understanding Hydrophobic Effects: Insights from Water Density Fluctuations. Annu. Rev. Condens. Matter. Phys. 2022, 13, 303–324. [Google Scholar] [CrossRef]
  7. Zhang, Q.-L.; Yang, R.-Y.; Wang, C.-L.; Hu, J. Ultrafast active water pump driven by terahertz electric fields. Phys. Rev. Fluids 2022, 7, 114202. [Google Scholar] [CrossRef]
  8. Zhong, J.; Zhu, C.; Li, L.; Richmond, G.L.; Francisco, J.S.; Zeng, X.C. Interaction of SO2 with the Surface of a Water Nanodroplet. J. Am. Chem. Soc. 2017, 139, 17168–17174. [Google Scholar] [CrossRef]
  9. Zhong, J.; Wang, C.; Zeng, X.C.; Francisco, J.S. Heterogeneous Reactions of SO3 on Ice: An Overlooked Sink for SO3 Depletion. J. Am. Chem. Soc. 2020, 142, 2150–2154. [Google Scholar] [CrossRef]
  10. Chandler, D. Interfaces and the driving force of hydrophobic assembly. Nature 2005, 437, 640–647. [Google Scholar] [CrossRef]
  11. Zhao, L.; Wang, C.L.; Liu, J.; Wen, B.H.; Tu, Y.S.; Wang, Z.W.; Fang, H.P. Reversible State Transition in Nanoconfined Aqueous Solutions. Phys. Rev. Lett. 2014, 112, 078301. [Google Scholar] [CrossRef]
  12. Prestipino, S.; Laio, A.; Tosatti, E. Systematic Improvement of Classical Nucleation Theory. Phys. Rev. Lett. 2012, 108, 225701. [Google Scholar] [CrossRef]
  13. Zhao, L.; Wang, C.; Fang, H.; Tu, Y. The Gibbs-free-energy landscape for the solute association in nanoconfined aqueous solutions. Nucl. Sci. Tech. 2015, 26, 030504. [Google Scholar]
  14. Yang, J.J.; Meng, S.; Xu, L.F.; Wang, E.G. Water adsorption on hydroxylated silica surfaces studied using the density functional theory. Phys. Rev. B 2005, 71, 035413. [Google Scholar] [CrossRef]
  15. Hu, X.L.; Michaelides, A. Water on the hydroxylated (001) surface of kaolinite: From monomer adsorption to a flat 2D wetting layer. Surf. Sci. 2008, 602, 960–974. [Google Scholar] [CrossRef]
  16. Michaelides, A.; Ranea, V.A.; de Andres, P.L.; King, D.A. General Model for Water Monomer Adsorption on Close-Packed Transition and Noble Metal Surfaces. Phys. Rev. Lett. 2003, 90, 216102. [Google Scholar] [CrossRef]
  17. Wang, C.L.; Zhao, L.; Zhang, D.H.; Chen, J.G.; Shi, G.S.; Fang, H.P. Upright or Flat Orientations of the Ethanol Molecules on Surface with Charge Dipoles and the Implication on the Wetting Behavior. J. Phys. Chem. C 2014, 118, 1873. [Google Scholar] [CrossRef]
  18. Liu, X.; Pang, H.; Liu, X.; Li, Q.; Zhang, N.; Mao, L.; Qiu, M.; Hu, B.; Yang, H.; Wang, X. Orderly porous covalent organic frameworks-based materials: Superior adsorbents for pollutants removal from aqueous solutions. Innovation 2021, 2, 100076. [Google Scholar] [CrossRef]
  19. Secchi, E.; Marbach, S.; Niguès, A.; Stein, D.; Siria, A.; Bocquet, L. Massive radius-dependent flow slippage in carbon nanotubes. Nature 2016, 537, 210–213. [Google Scholar] [CrossRef]
  20. Wu, K.; Chen, Z.; Li, J.; Li, X.; Xu, J.; Dong, X. Wettability effect on nanoconfined water flow. Proc. Natl. Acad. Sci. USA 2017, 114, 3358–3363. [Google Scholar] [CrossRef]
  21. Limmer, D.T.; Willard, A.P.; Madden, P.; Chandler, D. Hydration of metal surfaces can be dynamically heterogeneous and hydrophobic. Proc. Natl. Acad. Sci. USA 2013, 110, 4200–4205. [Google Scholar] [CrossRef] [Green Version]
  22. Cong, S.; Liu, X.; Jiang, Y.; Zhang, W.; Zhao, Z. Surface Enhanced Raman Scattering Revealed by Interfacial Charge-Transfer Transitions. Innovation 2020, 1, 100051. [Google Scholar] [CrossRef] [PubMed]
  23. Dobson, C.M.; Sali, A.; Karplus, M. Protein folding: A perspective from theory and experiment. Angew. Chem. Int. Edit. 1998, 37, 868–893. [Google Scholar] [CrossRef]
  24. Kwon, O.-H.; Yoo, T.H.; Othon, C.M.; Van Deventer, J.A.; Tirrell, D.A.; Zewail, A.H. Hydration dynamics at fluorinated protein surfaces. Proc. Natl. Acad. Sci. USA 2010, 107, 17101–17106. [Google Scholar] [CrossRef]
  25. Roche, J.; Caro, J.A.; Norberto, D.R.; Barthe, P.; Roumestand, C.; Schlessman, J.L.; Garcia, A.E.; García-Moreno, E.B.; Royer, C.A. Cavities determine the pressure unfolding of proteins. Proc. Natl. Acad. Sci. USA 2012, 109, 6945–6950. [Google Scholar] [CrossRef]
  26. Lin, M.M.; Zewail, A.H. Hydrophobic forces and the length limit of foldable protein domains. Proc. Natl. Acad. Sci. USA 2012, 109, 9851–9856. [Google Scholar] [CrossRef]
  27. Zhu, C.; Gao, Y.; Li, H.; Meng, S.; Li, L.; Francisco, J.S.; Zeng, X.C. Characterizing hydrophobicity of amino acid side chains in a protein environment via measuring contact angle of a water nanodroplet on planar peptide network. Proc. Natl. Acad. Sci. USA 2016, 113, 12946–12951. [Google Scholar] [CrossRef]
  28. Dai, B.; Kang, S.-g.; Huynh, T.; Lei, H.; Castelli, M.; Hu, J.; Zhang, Y.; Zhou, R. Salts drive controllable multilayered upright assembly of amyloid-like peptides at mica/water interface. Proc. Natl. Acad. Sci. USA 2013, 110, 8543–8548. [Google Scholar] [CrossRef]
  29. Grimm, B.; Schornbaum, J.; Jasch, H.; Trukhina, O.; Wessendorf, F.; Hirsch, A.; Torres, T.; Guldi, D.M. Step-by-step self-assembled hybrids that feature control over energy and charge transfer. Proc. Natl. Acad. Sci. USA 2012, 109, 15565–15571. [Google Scholar] [CrossRef]
  30. Bai, J.; Angell, C.A.; Zeng, X.C. Guest-free monolayer clathrate and its coexistence with two-dimensional high-density ice. Proc. Natl. Acad. Sci. USA 2010, 107, 5718–5722. [Google Scholar] [CrossRef]
  31. Zangi, R.; Mark, A.E. Monolayer Ice. Phys. Rev. Lett. 2003, 91, 025502. [Google Scholar] [CrossRef] [Green Version]
  32. Fang, H.; Wan, R.; Gong, X.; Lu, H.; Li, S. Dynamics of single-file water chains inside nanoscale channels: Physics, biological significance and applications. J. Phys. D App. Phys. 2008, 41, 103002. [Google Scholar] [CrossRef]
  33. Wan, R.; Li, J.; Lu, H.; Fang, H. Controllable Water Channel Gating of Nanometer Dimensions. J. Am. Chem. Soc. 2005, 127, 7166–7170. [Google Scholar] [CrossRef]
  34. Bai, J.; Zeng, X.C. Polymorphism and polyamorphism in bilayer water confined to slit nanopore under high pressure. Proc. Natl. Acad. Sci. USA 2012, 109, 21240–21245. [Google Scholar] [CrossRef]
  35. Nair, R.R.; Wu, H.A.; Jayaram, P.N.; Grigorieva, I.V.; Geim, A.K. Unimpeded Permeation of Water Through Helium-Leak–Tight Graphene-Based Membranes. Science 2012, 335, 442–444. [Google Scholar] [CrossRef]
  36. Gong, X.J.; Li, J.Y.; Zhang, H.; Wan, R.Z.; Lu, H.J.; Wang, S.; Fang, H.P. Enhancement of Water Permeation across a Nanochannel by the Structure outside the Channel. Phys. Rev. Lett. 2008, 101, 257801. [Google Scholar] [CrossRef] [PubMed]
  37. Wan, R.Z.; Lu, H.J.; Li, J.Y.; Bao, J.D.; Hu, J.; Fang, H.P. Concerted orientation induced unidirectional water transport through nanochannels. Phys. Chem. Chem. Phys. 2009, 11, 9898–9902. [Google Scholar] [CrossRef]
  38. Tu, Y.S.; Xiu, P.; Wan, R.Z.; Hu, J.; Zhou, R.H.; Fang, H.P. Water-mediated signal multiplication with Y-shaped carbon nanotubes. Proc. Natl. Acad. Sci. USA 2009, 106, 18120–18124. [Google Scholar] [CrossRef]
  39. Tu, Y.S.; Zhou, R.H.; Fang, H.P. Signal transmission, conversion and multiplication by polar molecules confined in nanochannels. Nanoscale 2010, 2, 1976–1983. [Google Scholar] [CrossRef]
  40. Lu, H.; Li, J.; Gong, X.; Wan, R.; Zeng, L.; Fang, H. Water permeation and wavelike density distributions inside narrow nanochannels. Phys. Rev. B 2008, 77, 174115. [Google Scholar] [CrossRef]
  41. Zhou, X.Y.; Wang, C.L.; Wu, F.M.; Feng, M.; Li, J.Y.; Lu, H.J.; Zhou, R.H. The ice-like water monolayer near the wall makes inner water shells diffuse faster inside a charged nanotube. J. Chem. Phys. 2013, 138, 204710. [Google Scholar] [CrossRef] [PubMed]
  42. Lai, C.-Y.; Tang, T.-C.; Amadei, C.A.; Marsden, A.J.; Verdaguer, A.; Wilson, N.; Chiesa, M. A nanoscopic approach to studying evolution in graphene wettability. Carbon 2014, 80, 784–792. [Google Scholar] [CrossRef]
  43. Amadei, C.A.; Lai, C.-Y.; Heskes, D.; Chiesa, M. Time dependent wettability of graphite upon ambient exposure: The role of water adsorption. J. Chem. Phys. 2014, 141, 084709. [Google Scholar] [CrossRef] [PubMed]
  44. Amadei, C.A.; Tang, T.C.; Chiesa, M.; Santos, S. The aging of a surface and the evolution of conservative and dissipative nanoscale interactions. J. Chem. Phys. 2013, 139, 084708. [Google Scholar] [CrossRef]
  45. Lu, J.-Y.; Lai, C.-Y.; Almansoori, I.; Chiesa, M. The evolution in graphitic surface wettability with first-principles quantum simulations: The counterintuitive role of water. Phys. Chem. Chem. Phys. 2018, 20, 22636–22644. [Google Scholar] [CrossRef]
  46. Hakim, L.; Kurniawan, I.D.O.; Indahyanti, E.; Pradana, I.P. Molecular Dynamics Simulation of Wetting Behavior: Contact Angle Dependency on Water Potential Models. ICS Phys. Chem. 2021, 1, 10. [Google Scholar] [CrossRef]
  47. Khalkhali, M.; Kazemi, N.; Zhang, H.; Liu, Q. Wetting at the nanoscale: A molecular dynamics study. J. Chem. Phys. 2017, 146, 114704. [Google Scholar] [CrossRef]
  48. Hung, S.-W.; Shiomi, J. Dynamic wetting of nanodroplets on smooth and patterned graphene-coated surface. J. Phys. Chem. C 2018, 122, 8423–8429. [Google Scholar] [CrossRef]
  49. Ball, P. Water as an active constituent in cell biology. Chem. Rev. 2008, 108, 74–108. [Google Scholar] [CrossRef]
  50. Ewing, G.E. Ambient thin film water on insulator surfaces. Chem. Rev. 2006, 106, 1511–1526. [Google Scholar] [CrossRef]
  51. Verdaguer, A.; Sacha, G.M.; Bluhm, H.; Salmeron, M. Molecular structure of water at interfaces: Wetting at the nanometer scale. Chem. Rev. 2006, 106, 1478–1510. [Google Scholar] [CrossRef]
  52. Hummer, G.; Rasaiah, J.C.; Noworyta, J.P. Water conduction through the hydrophobic channel of a carbon nanotube. Nature 2001, 414, 188–190. [Google Scholar] [CrossRef]
  53. Koga, K.; Gao, G.T.; Tanaka, H.; Zeng, X.C. Formation of ordered ice nanotubes inside carbon nanotubes. Nature 2001, 412, 802–805. [Google Scholar] [CrossRef]
  54. Hu, J.; Xiao, X.D.; Ogletree, D.F.; Salmeron, M. Imaging The Condensation and Evaporation of Molecularly Thin-films of Water with Nanometer Resolution. Science 1995, 268, 267–269. [Google Scholar] [CrossRef]
  55. Björneholm, O.; Hansen, M.H.; Hodgson, A.; Liu, L.-M.; Limmer, D.T.; Michaelides, A.; Pedevilla, P.; Rossmeisl, J.; Shen, H.; Tocci, G.; et al. Water at Interfaces. Chem. Rev. 2016, 116, 7698–7726. [Google Scholar] [CrossRef]
  56. Odelius, M.; Bernasconi, M.; Parrinello, M. Two dimensional ice adsorbed on mica surface. Phys. Rev. Lett. 1997, 78, 2855–2858. [Google Scholar] [CrossRef]
  57. Miranda, P.B.; Xu, L.; Shen, Y.R.; Salmeron, M. Icelike water monolayer adsorbed on mica at room temperature. Phys. Rev. Lett. 1998, 81, 5876–5879. [Google Scholar] [CrossRef]
  58. Guo, J.; Meng, X.; Chen, J.; Peng, J.; Sheng, J.; Li, X.-Z.; Xu, L.; Shi, J.-R.; Wang, E.; Jiang, Y. Real-space imaging of interfacial water with submolecular resolution. Nat. Mater. 2014, 13, 184. [Google Scholar] [CrossRef]
  59. Algara-Siller, G.; Lehtinen, O.; Wang, F.C.; Nair, R.R.; Kaiser, U.; Wu, H.A.; Geim, A.K.; Grigorieva, I.V. Square ice in graphene nanocapillaries. Nature 2015, 519, 443. [Google Scholar] [CrossRef]
  60. Koga, K.; Zeng, X.C.; Tanaka, H. Freezing of Confined Water: A Bilayer Ice Phase in Hydrophobic Nanopores. Phys. Rev. Lett. 1997, 79, 5262–5265. [Google Scholar] [CrossRef]
  61. Giovambattista, N.; Rossky, P.J.; Debenedetti, P.G. Phase Transitions Induced by Nanoconfinement in Liquid Water. Phys. Rev. Lett. 2009, 102, 050603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Han, S.H.; Choi, M.Y.; Kumar, P.; Stanley, H.E. Phase transitions in confined water nanofilms. Nature Phys. 2010, 6, 685–689. [Google Scholar] [CrossRef]
  63. Kapil, V.; Schran, C.; Zen, A.; Chen, J.; Pickard, C.J.; Michaelides, A. The first-principles phase diagram of monolayer nanoconfined water. Nature 2022, 609, 512–516. [Google Scholar] [CrossRef] [PubMed]
  64. Pal, S.K.; Zewail, A.H. Dynamics of water in biological recognition. Chem. Rev. 2004, 104, 2099–2123. [Google Scholar] [CrossRef]
  65. Liou, Y.-C.; Tocilj, A.; Davies, P.L.; Jia, Z. Mimicry of ice structure by surface hydroxyls and water of a [beta]-helix antifreeze protein. Nature 2000, 406, 322–324. [Google Scholar] [CrossRef]
  66. Gu, W.; Helms, V. Tightly Connected Water Wires Facilitate Fast Proton Uptake at The Proton Entrance of Proton Pumping Proteins. J. Am. Chem. Soc. 2009, 131, 2080–2081. [Google Scholar] [CrossRef]
  67. Kasson, P.M.; Lindahl, E.; Pande, V.S. Water Ordering at Membrane Interfaces Controls Fusion Dynamics. J. Am. Chem. Soc. 2011, 133, 3812–3815. [Google Scholar] [CrossRef]
  68. Raschke, T.M. Water structure and interactions with protein surfaces. Curr. Opin. Struc. Biol. 2006, 16, 152–159. [Google Scholar] [CrossRef]
  69. Kimmel, G.A.; Petrik, N.G.; Dohnalek, Z.; Kay, B.D. Crystalline ice growth on Pt(111): Observation of a hydrophobic water monolayer. Phys. Rev. Lett. 2005, 95, 166102. [Google Scholar] [CrossRef]
  70. Wang, C.; Lu, H.; Wang, Z.; Xiu, P.; Zhou, B.; Zuo, G.; Wan, R.; Hu, J.; Fang, H. Stable liquid water droplet on a water monolayer formed at room temperature on ionic model substrates. Phys. Rev. Lett. 2009, 103, 137801. [Google Scholar] [CrossRef]
  71. Wang, C.L.; Yang, Y.Z.; Fang, H.P. Recent advances on “ordered water monolayer that does not completely wet water” at room temperature. Sci. China-Phys. Mech. Astron. 2014, 57, 802–809. [Google Scholar] [CrossRef]
  72. Qi, C.; Zhou, B.; Wang, C.; Zheng, Y.; Fang, H. A nonmonotonic dependence of the contact angles on the surface polarity for a model solid surface. Phys. Chem. Chem. Phys. 2017, 19, 6665–6670. [Google Scholar] [CrossRef]
  73. Shao, S.; Zhao, L.; Guo, P.; Wang, C. Ordered Water Monolayer on Ionic Model Substrates Studied by Molecular Dynamics Simulations. Nucl. Sci. Tech. 2014, 25, 020502. [Google Scholar]
  74. Wang, C.L.; Zhou, B.; Xiu, P.; Fang, H.P. Effect of Surface Morphology on the Ordered Water Layer at Room Temperature. J. Phys. Chem. C 2011, 115, 3018–3024. [Google Scholar] [CrossRef]
  75. Qu, M.; Zhou, B.; Wang, C. Molecular simulation study of the adhesion work for water droplets on water monolayer at room temperature. Chin. Phy. B 2021, 30, 106804. [Google Scholar] [CrossRef]
  76. Cheh, J.; Gao, Y.; Wang, C.; Zhao, H.; Fang, H. Ice or water: Thermal properties of monolayer water adsorbed on a substrate. J. Stat. Mech. Theory Exp. 2013, 2013, P06009. [Google Scholar] [CrossRef]
  77. Qi, C.; Lei, X.; Zhou, B.; Wang, C.; Zheng, Y. Temperature regulation of the contact angle of water droplets on the solid surfaces. J. Chem. Phys. 2019, 150, 234703. [Google Scholar] [CrossRef]
  78. Zhu, Z.; Guo, H.; Jiang, X.; Chen, Y.; Song, B.; Zhu, Y.; Zhuang, S. Reversible Hydrophobicity–Hydrophilicity Transition Modulated by Surface Curvature. J. Phys. Chem. Lett. 2018, 9, 2346–2352. [Google Scholar] [CrossRef]
  79. Wang, C.L.; Wen, B.H.; Tu, Y.S.; Wan, R.Z.; Fang, H.P. Friction Reduction at a Superhydrophilic Surface: Role of Ordered Water. J. Phys. Chem. C 2015, 119, 11679–11684. [Google Scholar] [CrossRef]
  80. Ma, P.; Liu, Y.; Sang, X.; Tan, J.; Ye, S.; Ma, L.; Tian, Y. Homogeneous interfacial water structure favors realizing a low-friction coefficient state. J. Colloid Interface Sci. 2022, 626, 324–333. [Google Scholar] [CrossRef]
  81. Wu, S.; He, F.; Xie, G.; Bian, Z.; Luo, J.; Wen, S. Black Phosphorus: Degradation Favors Lubrication. Nano Lett. 2018, 18, 5618–5627. [Google Scholar] [CrossRef] [PubMed]
  82. Chen, S.F.; Jiang, S.Y. A new avenue to nonfouling materials. Adv. Mater. 2008, 20, 335–338. [Google Scholar] [CrossRef]
  83. Fujishima, A.; Zhang, X.; Tryk, D.A. TiO2 photocatalysis and related surface phenomena. Surf. Sci. Rep. 2008, 63, 515–582. [Google Scholar] [CrossRef]
  84. Liu, K.S.; Yao, X.; Jiang, L. Recent developments in bio-inspired special wettability. Chem. Soc. Rev. 2010, 39, 3240–3255. [Google Scholar] [CrossRef] [PubMed]
  85. Sun, T.L.; Feng, L.; Gao, X.F.; Jiang, L. Bioinspired surfaces with special wettability. Acc. Chem. Res. 2005, 38, 644–652. [Google Scholar] [CrossRef]
  86. Briscoe, W.H.; Titmuss, S.; Tiberg, F.; Thomas, R.K.; McGillivray, D.J.; Klein, J. Boundary lubrication under water. Nature 2006, 444, 191–194. [Google Scholar] [CrossRef]
  87. Drelich, J.; Chibowski, E.; Meng, D.D.; Terpilowski, K. Hydrophilic and superhydrophilic surfaces and materials. Soft Matter 2011, 7, 9804–9828. [Google Scholar] [CrossRef]
  88. Qu, M.; Huang, G.; Liu, X.; Nie, X.; Qi, C.; Wang, H.; Hu, J.; Fang, H.; Gao, Y.; Liu, W.-T.; et al. Room temperature bilayer water structures on a rutile TiO2(110) surface: Hydrophobic or hydrophilic? Chem. Sci. 2022, 13, 10546–10554. [Google Scholar] [CrossRef]
  89. Wang, R.; Hashimoto, K.; Fujishima, A.; Chikuni, M.; Kojima, E.; Kitamura, A.; Shimohigoshi, M.; Watanabe, T. Photogeneration of highly amphiphilic TiO2 surfaces. Adv. Mater. 1998, 10, 135–138. [Google Scholar] [CrossRef]
  90. Wang, R.; Hashimoto, K.; Fujishima, A.; Chikuni, M.; Kojima, E.; Kitamura, A.; Shimohigoshi, M.; Watanabe, T. Light-induced amphiphilic surfaces. Nature 1997, 388, 431–432. [Google Scholar] [CrossRef]
  91. Ohler, B.; Langel, W. Molecular dynamics simulations on the interface between titanium dioxide and water droplets: A new model for the contact angle. J. Phys. Chem. C 2009, 113, 10189–10197. [Google Scholar] [CrossRef]
  92. Hennessy, D.C.; Pierce, M.; Chang, K.-C.; Takakusagi, S.; You, H.; Uosaki, K. Hydrophilicity transition of the clean rutile TiO2 (1 1 0) surface. Electrochim. Acta 2008, 53, 6173–6177. [Google Scholar] [CrossRef]
  93. Liu, K.; Cao, M.; Fujishima, A.; Jiang, L. Bio-Inspired Titanium Dioxide Materials with Special Wettability and Their Applications. Chem. Rev. 2014, 114, 10044. [Google Scholar] [CrossRef]
  94. Thompson, T.L.; Yates, J.T. Surface science studies of the photoactivation of TiO2 new photochemical processes. Chem. Rev. 2006, 106, 4428–4453. [Google Scholar] [CrossRef]
  95. Zubkov, T.; Stahl, D.; Thompson, T.L.; Panayotov, D.; Diwald, O.; Yates, J.T. Ultraviolet light-induced hydrophilicity effect on TiO2 (110)(1×1). Dominant role of the photooxidation of adsorbed hydrocarbons causing wetting by water droplets. J. Phys. Chem. B 2005, 109, 15454–15462. [Google Scholar] [CrossRef]
  96. Takeuchi, M.; Sakamoto, K.; Martra, G.; Coluccia, S.; Anpo, M. Mechanism of photoinduced superhydrophilicity on the TiO2 photocatalyst surface. J. Phys. Chem. B 2005, 109, 15422–15428. [Google Scholar] [CrossRef]
  97. Balajka, J.; Hines, M.A.; DeBenedetti, W.J.; Komora, M.; Pavelec, J.; Schmid, M.; Diebold, U. High-affinity adsorption leads to molecularly ordered interfaces on TiO2 in air and solution. Science 2018, 361, 786–789. [Google Scholar] [CrossRef]
  98. Lee, K.; Kim, Q.; An, S.; An, J.; Kim, J.; Kim, B.; Jhe, W. Superwetting of TiO2 by light-induced water-layer growth via delocalized surface electrons. Proc. Natl. Acad. Sci. USA 2014, 111, 5784. [Google Scholar] [CrossRef]
  99. Kimmel, G.A.; Baer, M.; Petrik, N.G.; VandeVondele, J.; Rousseau, R.; Mundy, C.J. Polarization-and azimuth-resolved infrared spectroscopy of water on TiO2 (110): Anisotropy and the hydrogen-bonding network. J. Phys. Chem. Lett. 2012, 3, 778–784. [Google Scholar] [CrossRef]
  100. Petrik, N.G.; Kimmel, G.A. Hydrogen bonding, HD exchange, and molecular mobility in thin water films on TiO2 (110). Phy. Rev. Lett. 2007, 99, 196103. [Google Scholar] [CrossRef]
  101. Matthiesen, J.; Hansen, J.o.; Wendt, S.; Lira, E.; Schaub, R.; laegsgaard, E.; Besenbacher, F.; Hammer, B. Formation and diffusion of water dimers on rutile TiO2 (110). Phys. Rev. Lett. 2009, 102, 226101. [Google Scholar] [CrossRef] [PubMed]
  102. Lee, J.; Sorescu, D.C.; Deng, X.; Jordan, K.D. Water chain formation on TiO2 (110). J. Phys. Chem. Lett. 2012, 4, 53–57. [Google Scholar] [CrossRef] [PubMed]
  103. Wang, Z.-T.; Wang, Y.-G.; Mu, R.; Yoon, Y.; Dahal, A.; Schenter, G.K.; Glezakou, V.-A.; Rousseau, R.; Lyubinetsky, I.; Dohnálek, Z. Probing equilibrium of molecular and deprotonated water on TiO2 (110). Proc. Natl. Acad. Sci. USA 2017, 114, 1801–1805. [Google Scholar] [CrossRef] [PubMed]
  104. Tan, S.; Feng, H.; Zheng, Q.; Cui, X.; Zhao, J.; Luo, Y.; Yang, J.; Wang, B.; Hou, J.G. Interfacial hydrogen-bonding dynamics in surface-facilitated dehydrogenation of water on TiO2 (110). J. Am. Chem. Soc. 2019, 142, 826–834. [Google Scholar] [CrossRef]
  105. Serrano, G.; Bonanni, B.; Di Giovannantonio, M.; Kosmala, T.; Schmid, M.; Diebold, U.; Di Carlo, A.; Cheng, J.; VandeVondele, J.; Wandelt, K. Molecular ordering at the interface between liquid water and rutile TiO2 (110). Adv. Mater. Interfaces 2015, 2, 1500246. [Google Scholar] [CrossRef]
  106. Harris, L.A.; Quong, A.A. Molecular Chemisorption as the Theoretically Preferred Pathway for Water Adsorption on Ideal Rutile TiO2(110). Phys. Rev. Lett. 2004, 93, 086105. [Google Scholar] [CrossRef]
  107. Israelachvili, J.N. Intermolecular and Surface Forces; Academic Press: Cambridge, MA, USA, 2011. [Google Scholar]
  108. Xu, Z.; Gao, Y.; Wang, C.L.; Fang, H.P. Nano-scale Hydrophilicity on Metal Surfaces at Room Temperature: Coupling Lattice Constants and Crystal Faces. J. Phys. Chem. C 2015, 119, 20409. [Google Scholar] [CrossRef]
  109. Rotenberg, B.; Patel, A.J.; Chandler, D. Molecular Explanation for Why Talc Surfaces Can Be Both Hydrophilic and Hydrophobic. J. Am. Chem. Soc. 2011, 133, 20521–20527. [Google Scholar] [CrossRef]
  110. Giese, R.F.; Costanzo, P.M.; van Oss, C.J. The surface free energies of talc and pyrophyllite. Phys. Chem. Miner. 1991, 17, 611–616. [Google Scholar] [CrossRef]
  111. Phan, A.; Ho, T.A.; Cole, D.R.; Striolo, A. Molecular Structure and Dynamics in Thin Water Films at Metal Oxide Surfaces: Magnesium, Aluminum, and Silicon Oxide Surfaces. J. Phys. Chem. C 2012, 116, 15962–15973. [Google Scholar] [CrossRef]
  112. Gong, H.; Qi, C.; Yang, J.; Chen, J.; Lei, X.; Zhao, L.; Wang, C. Stable water droplets on composite structures formed by embedded water into fully hydroxylated β-cristobalite silica. Chin. Phy. B 2021, 30, 010503. [Google Scholar] [CrossRef]
  113. Wang, C.L.; Li, J.Y.; Fang, H.P. Ordered water monolayer at room temperature. Rend. Lincei 2011, 22, 5–16. [Google Scholar] [CrossRef]
  114. Lützenkirchen, J.; Franks, G.V.; Plaschke, M.; Zimmermann, R.; Heberling, F.; Abdelmonem, A.; Darbha, G.K.; Schild, D.; Filby, A.; Eng, P.; et al. The surface chemistry of sapphire-c: A literature review and a study on various factors influencing its IEP. Adv. Colloid Interface Sci. 2017, 251, 1–25. [Google Scholar] [CrossRef]
  115. Lützenkirchen, J.; Zimmermann, R.; Preocanin, T.; Filby, A.; Kupcik, T.; Kuttner, D.; Abdelmonem, A.; Schild, D.; Rabung, T.; Plaschke, M.; et al. An attempt to explain bimodal behaviour of the sapphire c-plane electrolyte interface. Adv. Colloid Interface Sci. 2010, 157, 61–74. [Google Scholar] [CrossRef]
  116. Zhang, J.; Tan, J.; Pei, R.; Ye, S.; Luo, Y. Ordered Water Layer on the Macroscopically Hydrophobic Fluorinated Polymer Surface and Its Ultrafast Vibrational Dynamics. J. Am. Chem. Soc. 2021, 143, 13074–13081. [Google Scholar] [CrossRef]
  117. Love, J.C.; Estroff, L.A.; Kriebel, J.K.; Nuzzo, R.G.; Whitesides, G.M. Self-assembled monolayers of thiolates on metals as a form of nanotechnology. Chem. Rev. 2005, 105, 1103–1169. [Google Scholar] [CrossRef]
  118. Major, R.C.; Houston, J.E.; McGrath, M.J.; Siepmann, J.I.; Zhu, X.Y. Viscous water meniscus under nanoconfinement. Phys. Rev. Lett. 2006, 96, 177803. [Google Scholar] [CrossRef]
  119. Lahann, J.; Mitragotri, S.; Tran, T.-N.; Kaido, H.; Sundaram, J.; Choi, I.S.; Hoffer, S.; Somorjai, G.A.; Langer, R. A reversibly switching surface. Science 2003, 299, 371–374. [Google Scholar] [CrossRef]
  120. Chen, J.Y.; Ratera, I.; Park, J.Y.; Salmeron, M. Velocity dependence of friction and hydrogen bonding effects. Phys. Rev. Lett. 2006, 96, 236102. [Google Scholar] [CrossRef]
  121. Pei, Y.; Ma, J. Electric field induced switching behaviors of monolayer-modified silicon surfaces: Surface designs and molecular dynamics simulations. J. Am. Chem. Soc. 2005, 127, 6802–6813. [Google Scholar] [CrossRef]
  122. Lee, S.-W.; Laibinis, P.E. Directed movement of liquids on patterned surfaces using noncovalent molecular adsorption. J. Am. Chem. Soc. 2000, 122, 5395–5396. [Google Scholar] [CrossRef]
  123. Ferguson, M.K.; Lohr, J.R.; Day, B.S.; Morris, J.R. Influence of buried hydrogen-bonding groups within monolayer films on gas-surface energy exchange and accommodation. Phys. Rev. Lett. 2004, 92, 073201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Konek, C.T.; Musorrafiti, M.J.; Al-Abadleh, H.A.; Bertin, P.A.; Nguyen, S.T.; Geiger, F.M. Interfacial acidities, charge densities, potentials, and energies of carboxylic acid-functionalized silica/water interfaces determined by second harmonic generation. J. Am. Chem. Soc. 2004, 126, 11754–11755. [Google Scholar] [CrossRef] [PubMed]
  125. Samanta, D.; Sarkar, A. Immobilization of bio-macromolecules on self-assembled monolayers: Methods and sensor applications. Chem. Soc. Rev. 2011, 40, 2567–2592. [Google Scholar] [CrossRef] [PubMed]
  126. Tu, A.; Kwag, H.R.; Barnette, A.L.; Kim, S.H. Water adsorption isotherms on CH3-, OH-, and COOH-terminated organic surfaces at ambient conditions measured with PM-RAIRS. Langmuir 2012, 28, 15263–15269. [Google Scholar] [CrossRef]
  127. Arima, Y.; Iwata, H. Effects of surface functional groups on protein adsorption and subsequent cell adhesion using self-assembled monolayers. J. Mater. Chem. 2007, 17, 4079–4087. [Google Scholar] [CrossRef]
  128. Choi, S.; Yang, Y.; Chae, J. Surface plasmon resonance protein sensor using Vroman effect. Biosens. Bioelectron. 2008, 24, 893–899. [Google Scholar] [CrossRef]
  129. Ohnuki, H.; Izumi, M.; Lenfant, S.; Guerin, D.; Imakubo, T.; Vuillaume, D. Deposition of TTF derivative on carboxyl terminated self-assembled monolayers. Appl. Surf. Sci. 2005, 246, 392–396. [Google Scholar] [CrossRef]
  130. James, M.; Darwish, T.A.; Ciampi, S.; Sylvester, S.O.; Zhang, Z.; Ng, A.; Gooding, J.J.; Hanley, T.L. Nanoscale condensation of water on self-assembled monolayers. Soft Matter 2011, 7, 5309–5318. [Google Scholar] [CrossRef]
  131. Guo, P.; Tu, Y.S.; Yang, J.R.; Wang, C.L.; Sheng, N.; Fang, H.P. Water-COOH Composite Structure with Enhanced Hydrophobicity Formed by Water Molecules Embedded into Carboxyl-Terminated Self-Assembled Monolayers. Phys. Rev. Lett. 2015, 115, 186101. [Google Scholar] [CrossRef]
  132. Wang, Y.; Duan, Z.; Fan, D. An Ion Diffusion Method for Visualising a Solid-like Water Nanofilm. Sci. Rep. 2013, 3, 3505. [Google Scholar] [CrossRef]
  133. Mao, D.; Wang, X.; Wu, Y.; Gu, Z.; Wang, C.; Tu, Y. Unexpected hydrophobicity on self-assembled monolayers terminated with two hydrophilic hydroxyl groups. Nanoscale 2021, 13, 19604–19609. [Google Scholar] [CrossRef]
  134. Massimiliano, S.; Spaldin, N.A. Origin of the dielectric dead layer in nanoscale capacitors. Nature 2006, 443, 679–682. [Google Scholar]
  135. Kim, Y.T.; Ito, Y.; Tadai, K.; Mitani, T.; Kim, U.-S.; Kim, H.-S.; Cho, B.-W. Drastic change of electric double layer capacitance by surface functionalization of carbon nanotubes. Appl. Phys. Lett. 2005, 87, 234106. [Google Scholar] [CrossRef]
  136. Ahmad, M.; Gu, W.; Geyer, T.; Helms, V. Adhesive Water Networks Facilitate Binding of Hydrophilic Protein Interfaces. Nat. Comm. 2011, 2, 1–7. [Google Scholar] [CrossRef]
  137. Wu, K.; Qi, C.; Zhu, Z.; Wang, C.; Song, B.; Chang, C. Terahertz Wave Accelerates DNA Unwinding: A Molecular Dynamics Simulation Study. J. Phys. Chem. Lett. 2020, 11, 7002–7008. [Google Scholar] [CrossRef]
  138. Szymczyk, A.; Fatin-Rouge, N.; Fievet, P.; Ramseyer, C.; Vidonne, A. Identification of dielectric effects in nanofiltration of metallic salts. J. Membr. Sci. 2007, 287, 102–110. [Google Scholar] [CrossRef]
  139. Zhu, Z.; Chang, C.; Shu, Y.; Song, B. Transition to a Superpermeation Phase of Confined Water Induced by a Terahertz Electromagnetic Wave. J. Phys. Chem. Lett. 2020, 11, 256–262. [Google Scholar] [CrossRef]
  140. Bonthuis, D.J.; Gekle, S.; Netz, R.R. Dielectric profile of interfacial water and its effect on double-layer capacitance. Phys. Rev. Lett. 2011, 107, 166102. [Google Scholar] [CrossRef]
  141. Fumagalli, L.; Esfandiar, A.; Fabregas, R.; Hu, S.; Ares, P.; Janardanan, A.; Yang, Q.; Radha, B.; Taniguchi, T.; Watanabe, K.; et al. Anomalously low dielectric constant of confined water. Science 2018, 360, 1339–1342. [Google Scholar] [CrossRef]
  142. Sato, T.; Sasaki, T.; Ohnuki, J.; Umezawa, K.; Takano, M. Hydrophobic Surface Enhances Electrostatic Interaction in Water. Phys. Rev. Lett. 2018, 121, 206002. [Google Scholar] [CrossRef] [PubMed]
  143. Sarhangi, S.M.; Waskasi, M.M.; Hashemianzadeh, S.M.; Matyushov, D.V. Effective Dielectric Constant of Water at the Interface with Charged C60 Fullerenes. J. Phys. Chem. B 2019, 123, 3135–3143. [Google Scholar] [CrossRef] [PubMed]
  144. Schlaich, A.; Knapp, E.W.; Netz, R.R. Water Dielectric Effects in Planar Confinement. Phys. Rev. Lett. 2016, 117, 048001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Zhang, C.; Gygi, F.; Galli, G. Strongly Anisotropic Dielectric Relaxation of Water at the Nanoscale. J. Phys. Chem. Lett. 2013, 4, 2477–2481. [Google Scholar] [CrossRef]
  146. Qi, W.; Zhao, H. Hydrogen bond network in the hydration layer of the water confined in nanotubes increasing the dielectric constant parallel along the nanotube axis. J. Chem. Phys. 2015, 143, 631. [Google Scholar]
  147. Conway, B.E.; Bockris, J.O.M.; Ammar, I.A. The dielectric constant of the solution in the diffuse and Helmholtz double layers at a charged interface in aqueous solution. Trans. Faraday Soc. 1951, 47, 756–766. [Google Scholar] [CrossRef]
  148. Hubbard, J.; Onsager, L. Dielectric dispersion and dielectric friction in electrolyte solutions. I. J. Chem. Phys. 1977, 67, 4850. [Google Scholar] [CrossRef]
  149. Chandra, A. Static dielectric constant of aqueous electrolyte solutions: Is there any dynamic contribution? J. Chem. Phys. 2000, 113, 903–905. [Google Scholar] [CrossRef]
  150. Zhu, H.; Ghoufi, A.; Szymczyk, A.; Balannec, B.; Morineau, D. Anomalous Dielectric Behavior of Nanoconfined Electrolytic Solutions. Phys. Rev. Lett. 2012, 109, 107801. [Google Scholar] [CrossRef]
  151. De Luca, S.; Kannam, S.K.; Todd, B.D.; Frascoli, F.; Hansen, J.S.; Daivis, P.J. Effects of Confinement on the Dielectric Response of Water Extends up to Mesoscale Dimensions. Langmuir 2016, 32, 4765–4773. [Google Scholar] [CrossRef]
  152. Varghese, S.; Kannam, S.K.; Hansen, J.S.; Sathian, S.P. Effect of Hydrogen Bonds on the Dielectric Properties of Interfacial Water. Langmuir 2019, 35, 8159–8166. [Google Scholar] [CrossRef]
  153. Zhang, C. Note: On the dielectric constant of nanoconfined water. J. Chem. Phys. 2018, 148, 156101. [Google Scholar] [CrossRef]
  154. Bonthuis, D.J.; Gekle, S.; and Netz, R.R. Profile of the Static Permittivity Tensor of Water at Interfaces: Consequences for Capacitance, Hydration Interaction and Ion Adsorption. Langmuir 2012, 28, 7679–7694. [Google Scholar] [CrossRef]
  155. Dong, A.; Yan, L.; Sun, L.; Yan, S.; Shan, X.; Guo, Y.; Meng, S.; Lu, X. Identifying Few-Molecule Water Clusters with High Precision on Au(111) Surface. ACS Nano 2018, 12, 6452–6457. [Google Scholar] [CrossRef]
  156. Li, S.; Chen, Y.; Zhao, J.; Wang, C.; Wei, N. Atomic structure rising obvious thermal conductance difference at Pd-H2O interface: A molecular dynamics simulation. Nanoscale 2020, 12, 17870. [Google Scholar] [CrossRef]
  157. Argyris, D.; Ho, T.; Cole, D.R.; Striolo, A. Molecular Dynamics Studies of Interfacial Water at the Alumina Surface. J. Phys. Chem. C 2011, 115, 2038–2046. [Google Scholar] [CrossRef]
  158. Qi, C.; Zhu, Z.; Wang, C.; Zheng, Y. Anomalously Low Dielectric Constant of Ordered Interfacial Water. J. Phys. Chem. Lett. 2021, 12, 931–937. [Google Scholar] [CrossRef]
  159. Shi, G.; Shen, Y.; Liu, J.; Wang, C.; Wang, Y.; Song, B.; Hu, J.; Fang, H. Molecular-scale Hydrophilicity Induced by Solute: Molecular-thick Charged Pancakes of Aqueous Salt Solution on Hydrophobic Carbon-based Surfaces. Sci. Rep. 2014, 4, 6793. [Google Scholar] [CrossRef]
  160. Wang, M.; Zhang, P.; Liang, X.; Zhao, J.; Liu, Y.; Cao, Y.; Wang, H.; Chen, Y.; Zhang, Z.; Pan, F.; et al. Ultrafast seawater desalination with covalent organic framework membranes. Nat. Sustain. 2022, 5, 518–526. [Google Scholar] [CrossRef]
  161. Qiao, Y.-Q.; Gu, Y.; Meng, Y.-S.; Li, H.-X.; Zhang, B.-W.; Li, J.-Y. Fabrication of stable MWCNT bucky paper for solar-driven interfacial evaporation by coupling γ-ray irradiation with borate crosslinking. Nucl. Sci. Tech. 2021, 32, 135. [Google Scholar] [CrossRef]
  162. Li, W.; Lai, J.; Zu, Y.; Lai, P. Cartilage-inspired hydrogel lubrication strategy. Innovation 2022, 3, 100275. [Google Scholar] [CrossRef] [PubMed]
  163. Gonella, G.; Backus, E.H.G.; Nagata, Y.; Bonthuis, D.J.; Loche, P.; Schlaich, A.; Netz, R.R.; Kühnle, A.; McCrum, I.T.; Koper, M.T.M.; et al. Water at charged interfaces. Nat. Rev. Chem. 2021, 5, 466–485. [Google Scholar] [CrossRef]
  164. Wilkins, D.M.; Manolopoulos, D.E.; Pipolo, S.; Laage, D.; Hynes, J.T. Nuclear Quantum Effects in Water Reorientation and Hydrogen-Bond Dynamics. J. Phys. Chem. Lett. 2017, 8, 2602–2607. [Google Scholar] [CrossRef] [PubMed]
  165. Guo, J.; Lü, J.-T.; Feng, Y.; Chen, J.; Peng, J.; Lin, Z.; Meng, X.; Wang, Z.; Li, X.-Z.; Wang, E.-G.; et al. Nuclear quantum effects of hydrogen bonds probed by tip-enhanced inelastic electron tunneling. Science 2016, 352, 321–325. [Google Scholar] [CrossRef]
  166. Sun, Z.; Zheng, L.; Chen, M.; Klein, M.L.; Paesani, F.; Wu, X. Electron-Hole Theory of the Effect of Quantum Nuclei on the X-Ray Absorption Spectra of Liquid Water. Phys. Rev. Lett. 2018, 121, 137401. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) A stable water droplet coexists with a thin water monolayer absorbed on a modeled solid surface. (b) Structure of the model solid surface with a hexagonal, charged pattern. Red presents the positive-charged atoms, and the blue presents the negative-charged atoms. (Reprinted from Ref. [70], Copyright 2009 American Physical Society).
Figure 1. (a) A stable water droplet coexists with a thin water monolayer absorbed on a modeled solid surface. (b) Structure of the model solid surface with a hexagonal, charged pattern. Red presents the positive-charged atoms, and the blue presents the negative-charged atoms. (Reprinted from Ref. [70], Copyright 2009 American Physical Society).
Crystals 13 00263 g001
Figure 2. (a) Side-view snapshot of the rutile TiO2(110) solid with a water droplet (red and white balls) coexisting with the water bilayer (cyan and white balls). (b) Snapshots of the ordered water structures for 2 MLs of H2O adsorbed on rutile TiO2 (110). The atoms are color coded as follows: Ti (pink), O of TiO2 (cyan), O of H2O (first layer: red; second layer: magenta), and H (white). (Reprinted from Ref. [88], Copyright 2022 Royal Society of Chemistry).
Figure 2. (a) Side-view snapshot of the rutile TiO2(110) solid with a water droplet (red and white balls) coexisting with the water bilayer (cyan and white balls). (b) Snapshots of the ordered water structures for 2 MLs of H2O adsorbed on rutile TiO2 (110). The atoms are color coded as follows: Ti (pink), O of TiO2 (cyan), O of H2O (first layer: red; second layer: magenta), and H (white). (Reprinted from Ref. [88], Copyright 2022 Royal Society of Chemistry).
Crystals 13 00263 g002
Figure 3. Typical side view of a water droplet on an ordered water monolayer on (a) Pd(100), (b) Pt(100), and (d) Al(100) surfaces, and the spreading water film on some typical surfaces: (c) Pd(110), (e) Pd(111), and (f) Ni(100). (g) Snapshot of the ordered, rhombic water molecules (in green lines) forming between neighboring water molecules. (h) Probability distribution of angle φ between the x–y plane projection of one water molecule dipole orientation and x-axis for (100) the crystal face. (Reprinted from Ref. [107], Copyright 2015 American Chemical Society).
Figure 3. Typical side view of a water droplet on an ordered water monolayer on (a) Pd(100), (b) Pt(100), and (d) Al(100) surfaces, and the spreading water film on some typical surfaces: (c) Pd(110), (e) Pd(111), and (f) Ni(100). (g) Snapshot of the ordered, rhombic water molecules (in green lines) forming between neighboring water molecules. (h) Probability distribution of angle φ between the x–y plane projection of one water molecule dipole orientation and x-axis for (100) the crystal face. (Reprinted from Ref. [107], Copyright 2015 American Chemical Society).
Crystals 13 00263 g003
Figure 4. (a) Side-view snapshots of water droplets and sparse water molecules outside the droplet on COOH-SAM with a packing density of Σ = 4.00 nm−2, together with top and side view snapshots of water on COOH SAM (model atoms: gray; COOH groups: blue, purple, and white; water: red and white; embedded water: green and white; alkyl chains: blue lines in side view, but omitted for clear views in the top view; H-bonds: red, dashed lines). (b) Dependence of contact angle values θ of water droplets on packing density Σ of COOH-SAMs. We marked the mid-range in light blue. (Reprinted from Ref. [131], Copyright 2015 American Physical Society).
Figure 4. (a) Side-view snapshots of water droplets and sparse water molecules outside the droplet on COOH-SAM with a packing density of Σ = 4.00 nm−2, together with top and side view snapshots of water on COOH SAM (model atoms: gray; COOH groups: blue, purple, and white; water: red and white; embedded water: green and white; alkyl chains: blue lines in side view, but omitted for clear views in the top view; H-bonds: red, dashed lines). (b) Dependence of contact angle values θ of water droplets on packing density Σ of COOH-SAMs. We marked the mid-range in light blue. (Reprinted from Ref. [131], Copyright 2015 American Physical Society).
Crystals 13 00263 g004
Figure 5. (a) Side-view snapshot of a water droplet with large contact angle of 82° on (OH)2-SAM with two hydroxyl (OH) groups at a packing density of Σ = 4.5 nm−2 (model atoms, gray; hydroxyl groups: purple and white; water: red and white; and alkyl chains: cyan and white). (b) Top-view snapshot (top) of the subfigure (bottom) with an enlarged region where a hexagonal-like H-bonding structure on (OH)2-SAM appears. (c) Contact angle θ of the water droplets on (OH)2-SAMs versus packing density Σ. (Reprinted from Ref. [133], Copyright 2021 Royal Society of Chemistry).
Figure 5. (a) Side-view snapshot of a water droplet with large contact angle of 82° on (OH)2-SAM with two hydroxyl (OH) groups at a packing density of Σ = 4.5 nm−2 (model atoms, gray; hydroxyl groups: purple and white; water: red and white; and alkyl chains: cyan and white). (b) Top-view snapshot (top) of the subfigure (bottom) with an enlarged region where a hexagonal-like H-bonding structure on (OH)2-SAM appears. (c) Contact angle θ of the water droplets on (OH)2-SAMs versus packing density Σ. (Reprinted from Ref. [133], Copyright 2021 Royal Society of Chemistry).
Crystals 13 00263 g005
Figure 6. Anomalously low, parallel, dielectric constant of interfacial water at the fcc (100) hydrophilic surface: (a) schematic representation of the confined water between two fcc (100) or (111) sheets separated by a distance of 10.0 nm in the z direction. (b) Top view of the lattice arrangement of the fcc (100) (left) and (111) (right) surface models, and (c) parallel permittivity of the dielectric constant of interfacial water ε Interfacial (equal to εxx and εyy) at the fcc (100) and (111) surfaces versus the surface-water interaction f (the orange, dashed line shows the constant for bulk water). (Reprinted from Ref. [158], Copyright 2021 American Chemical Society).
Figure 6. Anomalously low, parallel, dielectric constant of interfacial water at the fcc (100) hydrophilic surface: (a) schematic representation of the confined water between two fcc (100) or (111) sheets separated by a distance of 10.0 nm in the z direction. (b) Top view of the lattice arrangement of the fcc (100) (left) and (111) (right) surface models, and (c) parallel permittivity of the dielectric constant of interfacial water ε Interfacial (equal to εxx and εyy) at the fcc (100) and (111) surfaces versus the surface-water interaction f (the orange, dashed line shows the constant for bulk water). (Reprinted from Ref. [158], Copyright 2021 American Chemical Society).
Crystals 13 00263 g006
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

Qi, C.; Ling, C.; Wang, C. Ordered/Disordered Structures of Water at Solid/Liquid Interfaces. Crystals 2023, 13, 263. https://doi.org/10.3390/cryst13020263

AMA Style

Qi C, Ling C, Wang C. Ordered/Disordered Structures of Water at Solid/Liquid Interfaces. Crystals. 2023; 13(2):263. https://doi.org/10.3390/cryst13020263

Chicago/Turabian Style

Qi, Chonghai, Cheng Ling, and Chunlei Wang. 2023. "Ordered/Disordered Structures of Water at Solid/Liquid Interfaces" Crystals 13, no. 2: 263. https://doi.org/10.3390/cryst13020263

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