Surface: A Molecular Dynamics Simulation Work - ACS Publications

Mar 14, 2017 - cloud droplet formation,1 water harvesting,2,3 thermal manage- ment,4,5 water .... 0.1 ps relaxation time.50 All MD simulations are per...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/JPCC

Heterogeneous Condensation of Water on the Mica (001) Surface: A Molecular Dynamics Simulation Work Xinwen Ou,⊥,†,‡,§,∥ Xiaofeng Wang,⊥,‡ Zhang Lin,§,∥ and Jingyuan Li*,†,‡ †

Institute of Quantitative Biology and Department of Physics, Zhejiang University, Hangzhou 310027, China CAS Key Laboratory for Biomedical Effects of Nanomaterials and Nanosafety, Institute of High Energy Physics, Chinese Academy of Sciences, Beijing 100049, China § Key Laboratory of Design and Assembly of Functional Nanostructures, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002, China ∥ School of Environment and Energy, South China University of Technology, Guangzhou 510006, China ‡

S Supporting Information *

ABSTRACT: Water vapor condensation on the solid surface is ubiquitous in nature and has considerable importance in industrial applications. In this work, molecular dynamics simulation is used to investigate the kinetic process of water condensation on the mica surface as well as the properties of adsorbed water. The water molecules tend to spread on the mica surface and develop a water film comprised of two distinct adlayers. The growth of the first water adlayer is inhomogeneous due to the discrepancy of adsorption tendency to different surface sites. Interestingly, the second water adlayer begins to emerge far before the surface sites of mica are completely occupied. Our observations resemble the condensation process described by the Stranski−Krastanov growth model. These findings exhibit the dependence of water condensation on the surface properties.



INTRODUCTION Water vapor condensation on solid surfaces is of fundamental importance in various processes in nature and industry, e.g., cloud droplet formation,1 water harvesting,2,3 thermal management,4,5 water desalination,6 and antifogging coatings.7,8 The study about water condensation behavior and its influencing factors (e.g., surface properties, relative vapor pressures) will be helpful to understand these phenomena. So far, there are three primary growth models describing the process of water condensation on the solid surfaces, i.e., Volmer−Weber model (island formation), Frank−van der Merwe (FM) model (layer-by-layer growth), and Stranski−Krastanov (SK) model (layer-plus-island growth).9 Water condensation on a hydrophilic surface is usually characterized by the FM or SK model where a complete liquid film is initially formed. The formation of the first water adlayer serves as the key step of water condensation,9,10 and the effect of surface properties on water condensation is mainly reflected in this step. For example, surface defect often serves as the nucleation center and facilitates the formation of water adlayers.11−13 On the other hand, the formation process of water adlayers on a smooth (defect-free) surface appears to be much more complicated, and the understanding about this process is still limited.14,15 Mica substrate is widely used to study the properties of interfacial water and the formation process of a water adlayer, i.e., water condensation. A variety of features about structure © 2017 American Chemical Society

and dynamic behaviors of water in the vicinity of the mica (001) surface have been characterized by experimental studies using spectroscopic methods,16,17 X-ray reflectivity (XRR),18,19 surface force apparatus (SFA),20,21 scanning tunneling microscopy (STM),22 and atomic force microscopy (AFM).23−28 Cantrell et al. studied the absorption isotherm of water on the mica surface and found the adsorption enthalpy is related to the surface water coverage.16 Raviv et al. found that water molecules confined between two mica surfaces retain a shear fluidity characteristic of bulk liquid even under extreme confinement (D < 1.0 nm).20 Miranda et al. observed a stable ice-like ordered structure on the mica surface at room temperature,17 and such ice-like interfacial water structure facilitates the self-assembly of peptides on the mica surface.29,30 These studies have effectively characterized the properties of water adlayers. Nevertheless, detailed information about water condensation, e.g., growth process and the underlying molecular mechanism, is not unambiguous. Molecular dynamics (MD) and Monte Carlo (MC) simulation have been used to study the structure and dynamic properties of water adjacent to the mica substrate.31−38 Park et al. investigated the distribution of water adsorbed on the mica Received: January 26, 2017 Revised: March 14, 2017 Published: March 14, 2017 6813

DOI: 10.1021/acs.jpcc.7b00855 J. Phys. Chem. C 2017, 121, 6813−6819

Article

The Journal of Physical Chemistry C (001) surface and proposed strong lateral ordering of the hydration water.33 Moreover, the distribution of interfacial water on the mica surface was found to depend on relative vapor pressures.34 Wang et al. studied the diffusion of water adsorbed on the mica surface under different surface coverages.36 Leng et al. investigated the shear dynamics of hydration water between two mica surfaces and revealed that the persistent fluidity of the hydration layers is due to the fast rotational and translational dynamics of water molecules.37 These studies have provided molecular level understandings about the structure, dynamics, and energetics of water films on the mica surface. However, a limited attempt has been made to study the kinetic process of water condensation on the mica surface. In this work we use MD simulations to investigate the condensation process of water vapor on the mica surface, as well as the properties of adsorbed water. In the beginning, water molecules quickly approach onto the mica surface due to the strong interaction with the substrate, and the adsorbed water molecules tend to spread on the surface. The condensation process slows down as the water coverage increases. The resulting water film develops layering structure with two distinct adlayers. The growth of the first water adlayer is inhomogeneous due to different adsorption affinities to various surface sites of mica. More interestingly, the second water adlayer begins to emerge even when the surface sites are not completely occupied. In other words, the water molecules adsorbed in the preferential binding sites can recruit other water molecules, leading to the emergence of a second adlayer before the completion of the first adlayer. This phenomenon resembles the condensation process described by the SK growth model. Hence, our study suggests a molecular mechanism of the SK growth model; i.e., the formation of the first water adlayer resulted from the strong interaction with the substrate and the following island growth due to the inhomogeneous growth of the first adlayer. The water film becomes more ordered upon the coverage of graphene, which could be related to the enhanced hydrogen bond network of adsorbed water molecules.

Figure 1. (a) Initial configuration of the condensation system with the arrow denoting the deposition direction of water molecules. (b) The representative configuration of the system with ∼300 water molecules deposited on the upper surface of the mica substrate.

The system undergoes a standard conjugate gradient minimization, 1 ns water equilibrium, and 1600 ns production run in NVT (constant number of atoms, volume, and temperature) ensemble at 300 K, where the number of water molecules adsorbed on the upper surface continuously increases. The relative humidity of the system is 84.1%, comparable to the experimental observation that the second water adlayer appears with high relative humidity of ∼90%.16,27 Graphene has been extensively used as an ultrathin coating to enable high-resolution imaging of water film on the mica substrate. The impact of graphene coating to the properties of water film is also studied in this work. Ten systems are constructed with various numbers of adsorbed water molecules on the mica surface, i.e., water film containing 150, 200, 230, 250, 275, 300, 320, 350, 375, and 400 water molecules separately, and the graphene sheet with the size of 6.03 × 3.27 nm2 is placed upon the water film. For each system, a 3 ns NVT simulation is performed, and the last 2 ns trajectory is collected for analysis. The interaction of mica is described by the CLAYFF force field.41 Following previous studies, all atoms of the mica substrate are free to move during these simulations.42,43 The potential parameters for graphene are taken from the CHARMM27 force field.44 For water molecules, the TIP4P/ 2005 model is chosen,45 which is adequate to reproduce the properties of both the liquid and gas phase.46−48 A typical 1.2 nm cutoff distance is used to calculate the short-range electrostatic interaction as well as van der Waals interaction. The particle mesh Ewald method (PME) with a correction term for slab geometry is used to compute long-range electrostatic interactions.49 The system temperature is con-



SYSTEM AND METHOD The monoclinic C2/c 2M1-muscovite mica with a chemical formula KAl2(Si3Al)O10(OH)2 is a typical silicate consisting of aluminosilicate layers. One mica layer consists of two tetrahedral silicon sheets that are connected by one octahedral aluminum sheet, forming a TOT (tetrahedral−octahedral− tetrahedral) layer structure. Due to partial substitution of Al for Si in the tetrahedral sheets, the mica layer has a net negative structural charge which is neutralized by potassium ions. The OH groups of the octahedral sheet are parallel to the sheet plane. In our simulation, the unit cell parameters of mica are taken from the high-resolution X-ray reflectivity study.39 All tetrahedral Al atoms and potassium ions are arranged in an ordered fashion following an approach of Wang et al.36 As shown in Figure 1, the mica substrate consists of 12 × 4 × 1 crystallographic unit cells, corresponding to a dimension of 62.25 × 36.03 × 20.06 Å3. An aqueous region with ∼4100 water molecules is placed below the substrate, serving as a water reservoir with the thickness of ∼60 Å. The water molecules can evaporate freely from the reservoir and deposit on the upper surface of the mica substrate by the periodic boundary condition (PBC).40 The height of the simulation box is set to 200 Å. 6814

DOI: 10.1021/acs.jpcc.7b00855 J. Phys. Chem. C 2017, 121, 6813−6819

Article

The Journal of Physical Chemistry C trolled using the velocity-rescaled Berendsen thermostat with a 0.1 ps relaxation time.50 All MD simulations are performed using a time step of 1.0 fs with GROMACS 4.6.5 package.51



RESULTS AND DISCUSSION Water Film Growth and Properties of Adsorbed Water. We first investigate the kinetic process of water condensation on the mica (001) surface. The number of water molecules adsorbed on the upper surface of the mica substrate (N) is calculated (Figure 2). In the beginning, water molecules

Figure 3. Water density profiles for systems with a various number of water molecules adsorbed on the mica surface. The average position of surface bridging oxygen atoms is defined as the reference, i.e., z = 0.

bulk value (0.033/Å3), illustrating a well-structured hydration layer on the mica surface. The first three peaks (P1, P2, P3) correspond to water molecules directly adsorbed on the surface sites of mica (Figure 4). For peak P1, water molecules were centered over the

Figure 2. Time evolution of the total number of water molecules (N) deposited onto the mica surface. The two red lines are the fitted lines of the early stage (0−300 ns) and final stage (500−1600 ns).

quickly approach onto the mica surface due to strong interaction with the substrate and sufficiently spread out. For the first 300 ns, adsorbed water molecules accumulate linearly to N = 230 and tend to develop a water film. As the water coverage increases, the growth process begins to slow down. At time t = 500 ns, the adsorbed water molecules increase to N = 300. Thereafter, the water film grows in a much slower linear manner. At time t = 1600 ns, the adsorbed water molecules increase to N = 430, and the growth rate is about one-seventh of the first 300 ns. In other words, there are three distinct stages in the water condensation process (stage 1: N ≤ 230; stage 2:230 < N ≤ 300; stage 3: N > 300). During the water condensation process, two distinct water adlayers are formed sequentially. It should be noted that at stage 2 the second water adlayer has started to grow before the completion of the first water adlayer, resulting in an overlap of the growth of these two water adlayers which will be discussed later. Several representative structures of water film in these three stages are then studied. The water density profiles perpendicular to the surface in the cases of N = 150, 230, 300, and 400 are calculated separately. As demonstrated in Figure 3, there are four peaks (denoted as P1, P2, P3, and P4) that form and develop during the growth of water film, and the positions are z = 1.7, 2.7, 3.4, and 5.9 Å, respectively. The first three peaks constitute the first water adlayer (L1: 1.1 Å < z < 4.6 Å) with a thickness of 3.5 Å. The fourth peak (P4) is attributed to the second water adlayer (L2: 4.6 Å < z < 8.1 Å). The thickness of each adlayer is very close to that of a single ice layer (ice Ih, ∼3.7 Å), in agreement with the result of the previous AFM experiment.27 When N ≤ 230, only the first water adlayer L1 can be observed. As the water coverage increases, the growth of the second water adlayer L2 prevails. It should be noted that the water density at P2 (0.098/Å3) can be up to thrice of the

Figure 4. Snapshots of water molecules adsorbed in three surface sites of mica, from S1 to S3. Both top view and side view are presented.

siloxane ring (denoted as surface site S1). Peak P2 corresponds to the water molecules above the surface bridging oxygen (i.e., site S2). Peak P3 corresponds to the water molecules above the potassium ions (i.e., site S3). The water molecules in the sites S2 and S3 form the solvation shells of surface potassium ions. These binding sites were also identified in the ab initio simulations of Odelius et al.32 In addition, these water molecules have distinct dipole orientations according to their binding sites. The angle between the water dipole and surface normal, φD, was calculated (Figure S1). For water molecules in site S1, the dipole orientation φD is around 180°. The water dipole in site S2 also points downward, with φD ranging from 130 to 160°, while for the site S3, the water dipole adopts upward orientation; i.e., φD distributes between 20° and 90°. This indicates the interaction of water molecules with various surface sites may be different from each other. The anisotropic nature of water adsorption results in the inhomogeneous growth of the first adlayer. In the beginning (N = 150), water molecules tend to adsorb in binding site S2, and the water molecules mainly distribute in peak P2. Subsequently, the peaks P1 and P3 increase (N = 230). Interestingly, as the adsorption proceeds the peak P2 decreases (N = 300). It suggests the redistribution of water molecules within the first adlayer. The numbers of water molecules in the three binding sites are then calculated separately (Figure 5). The water molecules in the binding site S2 reach the maximum of N = 124 at time t = 230 ns. Thereafter, the water molecules in the site S2 gradually decrease to N = 100 at time t = 400 ns; i.e., there 6815

DOI: 10.1021/acs.jpcc.7b00855 J. Phys. Chem. C 2017, 121, 6813−6819

Article

The Journal of Physical Chemistry C

interact with other water molecules, resulting in the decrease of water adsorption rate. The dynamic behavior of interfacial water is also investigated. The method developed by Berne and co-workers is adopted, and the diffusion coefficient is evaluated as the parallel component of the diffusion tensor, Dxy52 Dxy = lim

τ→∞

⟨[Δx 2(τ ) + Δy 2 (τ )]⟩ 4τ⟨P(τ )⟩

(1)

where τ is time interval; P(τ) is the survival probability of water molecules in adlayers; and Δx2(τ) and Δy2(τ) are the mean square displacements in the x and y direction. At relatively low surface coverage (N ≤ 230) Dxy decreases as the water number increases. When the first water adlayer is almost complete (N = 230) Dxy decreases to 0.78 × 10−5 cm2/s. At the same time, the water molecules within the first adlayer have developed a lateral hydrogen bond network. Our observation is consistent with the previous study, wherein the organized hydrogen bond network at monolayer coverage induces the minimum of diffusion coefficient.43 At higher surface coverage (N > 230) Dxy increases, mainly attributed to relatively fast motion of water molecules in the second adlayer. To better understand the condensation process, we study the interaction energy for adsorbed water molecules during the growth of the water adlayer. The interaction energy between water molecules and mica, among the water molecules, is calculated separately (Figure 7). The change of these two

Figure 5. Time evolution of water molecules in the first adlayer. The blue line is the number of water molecules in the surface site S1; the red one is for site S2; and the green one is for site S3. The black line is the sum of the three sites, corresponding to the water number of the first water adlayer L1.

are 20% of the water molecules redistributed to the other binding sites. As mentioned above, during the early stage of condensation (N < 230) the water molecules directly adsorb to the substrate, forming the first adlayer. Interestingly, the second water adlayer begins to emerge even when some surface sites are still empty (Figure 6a). Moreover, the emergence of a second layer corresponds to the slowdown of water condensation in stage 2 (230 < N ≤ 300, Figure 2). At the end of stage 2, the number of water molecules in the first adlayer increases up to 230 and remains constant thereafter. In the third stage of condensation (N > 300), the number of water molecules in the second adlayer slowly increases. In short, the second water adlayer begins to grow before the surface sites are completely occupied. This scenario can be described by the SK model rather than the VW and FM growth model.9,10 The validity of the SK growth model in this system could be related to the discrepancy of water adsorption tendency to different surface sites. The adsorption affinity for water molecules in the binding sites S1 and S3 is noticeably weaker than that of site S2. When the preferential binding sites (S2) are completely occupied, a considerable portion of less favorable sites (S1 and S3) remain unoccupied. Subsequently, the water molecules adsorbed in binding site S2 begin to recruit other water molecules to adsorb above, and the second adlayer begins to emerge. It is worth mentioning that the newly deposited water molecules in the second adlayer can only

Figure 7. Average interaction energy of adsorbed water molecules (black squares). The interaction energy with mica (red dots) as well as interaction energy with other water molecules (blue triangles) are calculated separately.

Figure 6. (a) Number of water molecules in each adlayer during the condensation process. The black square is for water number of L1 corresponding to the left y axis, and the red dot is for water number of L2 corresponding to the right y axis. (b) The parallel component of the diffusion tensor of interfacial water. 6816

DOI: 10.1021/acs.jpcc.7b00855 J. Phys. Chem. C 2017, 121, 6813−6819

Article

The Journal of Physical Chemistry C

molecule and enhances the hydrogen bond networks of adlayers.

interaction energies exhibits opposite tendency as the water film grows: the interaction among water molecules gets enhanced, while the interaction between water molecules and the mica substrate becomes weaker. The average binding energy per water molecule31 for the water adlayers ranges from 60 to 70 kJ/mol, comparable to the cohesive energy of Ice Ih (∼65 kJ/mol),31 and the binding energy decreases with the growth of water film, in good agreement with the previous experiments which suggests hydration energy of mica decreases as the surface coverage increases.53 Interestingly, the tendency that the hydration energy of mica gets weaker is also reflected in the slowdown of water condensation. Graphene Coating Promotes the Layering Structure of Adsorbed Water. Graphene has been extensively used as a coating to visualize water adlayers on mica surfaces.22,25−28,54 However, the impacts of the graphene coating on the properties of water adlayers are not clear. We first study the impact on the structure of the water adlayer (Figure 8a). The water film



SUMMARY AND CONCLUSIONS In summary, the condensation process of water vapor on the mica surface as well as the properties of interfacial water are investigated. In the early stage the water condensation process is fast due to strong interaction with the substrate, and the process slows down as the water coverage increases. The adsorbed water molecules tend to spread on the mica surface and develop a water film comprised of two distinct adlayers, and the coating of graphene further enhances the layering structure of the water film. The growth of the first adlayer is heterogeneous due to different adsorption affinities to various surface sites. Interestingly, the second water adlayer begins to emerge even when some surface sites are still empty: the water molecules adsorbed in the preferential binding sites can effectively recruit other water molecules. The condensation process of water film on the mica surface resembles the process described by the SK model, i.e., the formation of the first adlayer followed by island growth. Hence, our results suggest the SK growth process may be related to the discrepancy of water adsorption tendency to different surface sites. Further studies about the condensation process of water vapor on other substrates are highly demanded to systematically understand the relationship between condensation behavior and surface properties.



ASSOCIATED CONTENT

S Supporting Information *

Figure 8. (a) Representative configuration of the system (N = 400) wherein the water film is covered by graphene and the vacuum slab is not represented. (b) The average number of hydrogen bonds per water molecule, and the systems in either the absence (red dots) or presence (black squares) of graphene are compared.

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.7b00855. The distributions of dipole orientation of interfacial water (Figure S1); the comparison of water density profiles of both systems (Figure S2); the distributions of dipole orientation of water adlayers with graphene coating (Figure S3); the self-diffusion coefficients of interfacial water with graphene coating (Figure S4) (PDF)

covered by graphene also has distinct layering structure, and the positions of the first three peaks are almost identical to the case without graphene. In addition, graphene coating induces the higher crests and deeper troughs in the density profiles and affects the relative distribution among these peaks (Figure S2). For example, in the case of N = 200, peak P3 eliminates, and peak P2 enhances. For N = 300, the sprouting second water adlayer is much weaker, and peak P3 gets stronger. For N = 400, the major part of peak P4 becomes more intensive, and its tail largely eliminates. Taken together, the water film becomes more ordered after graphene coating. The dipole orientation of water molecules in the first adlayer is similar to the case in the absence of graphene (Figure S3), while the distribution of dipole orientation of water molecules in the second adlayer is much narrower. It also suggests the water adlayer becomes more ordered. In both systems, the diffusion coefficient of adsorbed water is much smaller than the bulk water. And the decrement is more profound after graphene coating: the parallel component of the diffusion tensor can be as low as 0.41 × 10−5cm2/s (Figure S4). In addition, we study the hydrogen bond networks of water adlayer by calculating the average number of hydrogen bonds for water molecules.55 The hydrogen bonds with surface oxygen atoms of mica, as well as the hydrogen bonds among the water molecules, are taken into account (Figure 8b). The number of hydrogen bonds per water molecule increases as the water film grows in both systems. The coating of graphene further facilitates the hydrogen bond interaction of the adsorbed water



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Zhang Lin: 0000-0002-6600-2055 Jingyuan Li: 0000-0003-2926-1864 Author Contributions ⊥

Xinwen Ou and Xiaofeng Wang contributed equally to this work.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We thank Dr. Jicun Li for helpful discussions. This work is supported by the 100-talent project of Chinese Academy of Science, Ministry of Science and Technology (MOST) 973 program (2013CB933704), the National Natural Science Foundation of China (NSFC) grants (21473207), and the Special Program for Applied Research on Super Computation of the NSFC-Guangdong Joint Fund (the second phase). 6817

DOI: 10.1021/acs.jpcc.7b00855 J. Phys. Chem. C 2017, 121, 6813−6819

Article

The Journal of Physical Chemistry C



(25) Song, J.; Li, Q.; Wang, X.; Li, J.; Zhang, S.; Kjems, J.; Besenbacher, F.; Dong, M. Evidence of Stranski-Krastanov Growth at the Initial Stage of Atmospheric Water Condensation. Nat. Commun. 2014, 5, 4837. (26) Kim, J. S.; Choi, J. S.; Lee, M. J.; Park, B. H.; Bukhvalov, D.; Son, Y. W.; Yoon, D.; Cheong, H.; Yun, J. N.; Jung, Y.; et al. Between Scylla and Charybdis: Hydrophobic Graphene-Guided Water Diffusion on Hydrophilic Substrates. Sci. Rep. 2013, 3, 2309. (27) Xu, K.; Cao, P.; Heath, J. R. Graphene Visualizes the First Water Adlayers on Mica at Ambient Conditions. Science 2010, 329, 1188− 1191. (28) Cheng, M.; Wang, D.; Sun, Z.; Zhao, J.; Yang, R.; Wang, G.; Yang, W.; Xie, G.; Zhang, J.; Chen, P.; et al. A Route toward Digital Manipulation of Water Nanodroplets on Surfaces. ACS Nano 2014, 8, 3955−3960. (29) Kang, S. G.; Li, H.; Huynh, T.; Zhang, F.; Xia, Z.; Zhang, Y.; Zhou, R. Molecular Mechanism of Surface-Assisted Epitaxial SelfAssembly of Amyloid-Like Peptides. ACS Nano 2012, 6, 9276−9282. (30) Kang, S.-g.; Huynh, T.; Xia, Z.; Zhang, Y.; Fang, H.; Wei, G.; Zhou, R. Hydrophobic Interaction Drives Surface-Assisted Epitaxial Assembly of Amyloid-Like Peptides. J. Am. Chem. Soc. 2013, 135, 3150−3157. (31) Li, H.; Zeng, X. C. Two Dimensional Epitaxial Water Adlayer on Mica with Graphene Coating: An Ab Initio Molecular Dynamics Study. J. Chem. Theory Comput. 2012, 8, 3034−3043. (32) Odelius, M.; Bernasconi, M.; Parrinello, M. Two Dimensional Ice Adsorbed on Mica Surface. Phys. Rev. Lett. 1997, 78, 2855−2858. (33) Park, S. H.; Sposito, G. Structure of Water Adsorbed on a Mica Surface. Phys. Rev. Lett. 2002, 89, 085501. (34) Malani, A.; Ayappa, K. G. Adsorption Isotherms of Water on Mica: Redistribution and Film Growth. J. Phys. Chem. B 2009, 113, 1058−1067. (35) Sakuma, H.; Kawamura, K. Structure and Dynamics of Water on Muscovite Mica Surfaces. Geochim. Cosmochim. Acta 2009, 73, 4100− 4110. (36) Wang, J.; Kalinichev, A. G.; Kirkpatrick, R. J.; Cygan, R. T. Structure, Energetics, and Dynamics of Water Adsorbed on the Muscovite (001) Surface: A Molecular Dynamics Simulation. J. Phys. Chem. B 2005, 109, 15893−15905. (37) Leng, Y.; Cummings, P. T. Fluidity of Hydration Layers Nanoconfined between Mica Surfaces. Phys. Rev. Lett. 2005, 94, 026101. (38) Meleshyn, A. Two-Dimensional Ordering of Water Adsorbed on a Mica Surface at Room Temperature. J. Phys. Chem. C 2008, 112, 14495−14500. (39) Schlegel, M. L.; Nagy, K. L.; Fenter, P.; Cheng, L.; Sturchio, N. C.; Jacobsen, S. D. Cation Sorption on the Muscovite (001) Surface in Chloride Solutions Using High-Resolution X-Ray Reflectivity. Geochim. Cosmochim. Acta 2006, 70, 3549−3565. (40) Dolan, E. A.; Venable, R. M.; Pastor, R. W.; Brooks, B. R. Simulations of Membranes and Other Interfacial Systems Using P21 and Pc Periodic Boundary Conditions. Biophys. J. 2002, 82, 2317− 2325. (41) Cygan, R. T.; Liang, J.-J.; Kalinichev, A. G. Molecular Models of Hydroxide, Oxyhydroxide, and Clay Phases and the Development of a General Force Field. J. Phys. Chem. B 2004, 108, 1255−1266. (42) Malani, A.; Ayappa, K. G. Relaxation and Jump Dynamics of Water at the Mica Interface. J. Chem. Phys. 2012, 136, 194701. (43) Wang, J.; Kalinichev, A. G.; Kirkpatrick, R. J. Effects of Substrate Structure and Composition on the Structure, Dynamics, and Energetics of Water at Mineral Surfaces: A Molecular Dynamics Modeling Study. Geochim. Cosmochim. Acta 2006, 70, 562−582. (44) MacKerell, A. D.; Bashford, D.; Bellott, M.; Dunbrack, R. L.; Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; et al. All-Atom Empirical Potential for Molecular Modeling and Dynamics Studies of Proteins. J. Phys. Chem. B 1998, 102, 3586−3616. (45) Abascal, J. L.; Vega, C. A General Purpose Model for the Condensed Phases of Water: Tip4p/2005. J. Chem. Phys. 2005, 123, 234505.

REFERENCES

(1) Kulmala, M.; Pirjola, L.; Makela, J. M. Stable Sulphate Clusters as a Source of New Atmospheric Particles. Nature 2000, 404, 66−69. (2) Park, K. C.; Kim, P.; Grinthal, A.; He, N.; Fox, D.; Weaver, J. C.; Aizenberg, J. Condensation on Slippery Asymmetric Bumps. Nature 2016, 531, 78−82. (3) Lee, A.; Moon, M. W.; Lim, H.; Kim, W. D.; Kim, H. Y. Water Harvest Via Dewing. Langmuir 2012, 28, 10183−10191. (4) Rose, J. W. Condensation Heat Transfer Fundamentals. Chem. Eng. Res. Des. 1998, 76, 143−152. (5) Miljkovic, N.; Enright, R.; Wang, E. N. Effect of Droplet Morphology on Growth Dynamics and Heat Transfer During Condensation on Superhydrophobic Nanostructured Surfaces. ACS Nano 2012, 6, 1776−1785. (6) Khawaji, A. D.; Kutubkhanah, I. K.; Wie, J.-M. Advances in Seawater Desalination Technologies. Desalination 2008, 221, 47−69. (7) Introzzi, L.; Fuentes-Alventosa, J. M.; Cozzolino, C. A.; Trabattoni, S.; Tavazzi, S.; Bianchi, C. L.; Schiraldi, A.; Piergiovanni, L.; Farris, S. ″Wetting Enhancer″ Pullulan Coating for Antifog Packaging Applications. ACS Appl. Mater. Interfaces 2012, 4, 3692− 3700. (8) Di Mundo, R.; d’Agostino, R.; Palumbo, F. Long-Lasting Antifog Plasma Modification of Transparent Plastics. ACS Appl. Mater. Interfaces 2014, 6, 17059−17066. (9) Bauer, E. Phänomenologische Theorie Der Kristallabscheidung an Oberflächen. I. Z. Kristallogr. 1958, 110, 372−394. (10) Stranski, I. N.; Krastanow, L. Zur Theorie Der Orientierten Ausscheidung Von Ionenkristallen Aufeinander. Monatsh. Chem. 1937, 71, 351−364. (11) Li, Q.; Song, J.; Besenbacher, F.; Dong, M. Two-Dimensional Material Confined Water. Acc. Chem. Res. 2015, 48, 119−127. (12) Cailliez, F.; Stirnemann, G.; Boutin, A.; Demachy, I.; Fuchs, A. H. Does Water Condense in Hydrophobic Cavities? A Molecular Simulation Study of Hydration in Heterogeneous Nanopores. J. Phys. Chem. C 2008, 112, 10435−10445. (13) Trzpit, M.; Soulard, M.; Patarin, J.; Desbiens, N.; Cailliez, F.; Boutin, A.; Demachy, I.; Fuchs, A. H. The Effect of Local Defects on Water Adsorption in Silicalite-1 Zeolite: A Joint Experimental and Molecular Simulation Study. Langmuir 2007, 23, 10131−10139. (14) Shevkunov, S. V. Nucleation of Water Vapor in Microcracks on the Surface of B-Agi Aerosol Particles: 2. Thermodynamics of Nucleation. Colloid J. 2007, 69, 378−390. (15) Thiel, P. A.; Madey, T. E. The Interaction of Water with Solid Surfaces: Fundamental Aspects. Surf. Sci. Rep. 1987, 7, 211−385. (16) Cantrell; Ewing, G. E. Thin Film Water on Muscovite Mica. J. Phys. Chem. B 2001, 105, 5434−5439. (17) 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. (18) Cheng, L.; Fenter, P.; Nagy, K. L.; Schlegel, M. L.; Sturchio, N. C. Molecular-Scale Density Oscillations in Water Adjacent to a Mica Surface. Phys. Rev. Lett. 2001, 87, 156103. (19) Park, C.; Fenter, P. A.; Nagy, K. L.; Sturchio, N. C. Hydration and Distribution of Ions at the Mica-Water Interface. Phys. Rev. Lett. 2006, 97, 016101. (20) Raviv, U.; Klein, J. Fluidity of Bound Hydration Layers. Science 2002, 297, 1540−1543. (21) Raviv, U.; Laurat, P.; Klein, J. Fluidity of Water Confined to Subnanometre Films. Nature 2001, 413, 51−54. (22) He, K. T.; Wood, J. D.; Doidge, G. P.; Pop, E.; Lyding, J. W. Scanning Tunneling Microscopy Study and Nanomanipulation of Graphene-Coated Water on Mica. Nano Lett. 2012, 12, 2665−2672. (23) Spagnoli, C.; Loos, K.; Ulman, A.; Cowman, M. K. Imaging Structured Water and Bound Polysaccharide on Mica Surface at Ambient Temperature. J. Am. Chem. Soc. 2003, 125, 7124−7128. (24) 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. 6818

DOI: 10.1021/acs.jpcc.7b00855 J. Phys. Chem. C 2017, 121, 6813−6819

Article

The Journal of Physical Chemistry C (46) Vega, C.; Abascal, J. L. Simulating Water with Rigid NonPolarizable Models: A General Perspective. Phys. Chem. Chem. Phys. 2011, 13, 19663−19688. (47) Sakamaki, R.; Sum, A. K.; Narumi, T.; Yasuoka, K. Molecular Dynamics Simulations of Vapor/Liquid Coexistence Using the Nonpolarizable Water Models. J. Chem. Phys. 2011, 134, 124708. (48) Yang, C.; Jang, S.; Pak, Y. A Fully Atomistic Computer Simulation Study of Cold Denaturation of a Beta-Hairpin. Nat. Commun. 2014, 5, 5773. (49) Yeh, I.-C.; Berkowitz, M. L. Ewald Summation for Systems with Slab Geometry. J. Chem. Phys. 1999, 111, 3155. (50) Bussi, G.; Donadio, D.; Parrinello, M. Canonical Sampling through Velocity Rescaling. J. Chem. Phys. 2007, 126, 014101. (51) Hess, B.; Kutzner, C.; van der Spoel, D.; Lindahl, E. Gromacs 4: Algorithms for Highly Efficient, Load-Balanced, and Scalable Molecular Simulation. J. Chem. Theory Comput. 2008, 4, 435−447. (52) Liu, P.; Harder, E.; Berne, B. J. On the Calculation of Diffusion Coefficients in Confined Fluids and Interfaces with an Application to the Liquid-Vapor Interface of Water. J. Phys. Chem. B 2004, 108, 6595−6602. (53) Rakhmatkariev, G. Mechanism of Adsorption of Water Vapor by Muscovite: A Model Based on Adsorption Calorimetry. Clays Clay Miner. 2006, 54, 402−407. (54) Severin, N.; Lange, P.; Sokolov, I. M.; Rabe, J. P. Reversible Dewetting of a Molecularly Thin Fluid Water Film in a Soft GrapheneMica Slit Pore. Nano Lett. 2012, 12, 774−779. (55) Ou, X.; Li, J.; Lin, Z. Dynamic Behavior of Interfacial Water on Mg(OH)2 (001) Surface: A Molecular Dynamics Simulation Work. J. Phys. Chem. C 2014, 118, 29887−29895.

6819

DOI: 10.1021/acs.jpcc.7b00855 J. Phys. Chem. C 2017, 121, 6813−6819