Density Functional Theory Study of Water Molecule Adsorption on the

Jul 26, 2019 - (25−28) The thickness of the hydration layer on the silica surface was ... (101) surface and α-Bromolauric acid (CH3(CH2)9CHBrCOOH, ...
0 downloads 0 Views 3MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2019, 4, 12711−12718

http://pubs.acs.org/journal/acsodf

Density Functional Theory Study of Water Molecule Adsorption on the α‑Quartz (001) Surface with and without the Presence of Na+, Mg2+, and Ca2+ Chunfu Liu, Fanfei Min,* Lingyun Liu, and Jun Chen School of Materials Science and Engineering, Anhui University of Science and Technology, Huainan 232001, China

Downloaded via 46.148.127.34 on August 7, 2019 at 11:29:41 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: Adsorption of the single water molecule on the α-quartz (001) surface with and without the presence of Na+, Mg2+ and Ca2+ was analyzed utilizing the density functional theory method. Our results demonstrate that the optimal adsorption configuration of the single water molecule on the αquartz (001) surface lies in the bridge being configured with two formed hydrogen bonds. These were Os−Hw and Hs−Ow (s and w represent, respectively, surface and water molecules), while the main hydrogen bond is Hw−Os. Furthermore, the corresponding adsorption energy was ∼−72.60 kJ/mol. In this study, the presence of metal ions helped to deflect the spatial position of the water molecule, and the distance between Ow and Hs was altered significantly. Furthermore, the charge transfer between the interacting atoms increased in the presence of metal ions, wherein the effects of Ca2+ and Na+ proved to be significant compared to Mg2+. Finally, it emerged that metal ions interacted with the water molecule and were subsequently adsorbed on the α-quartz (001) surface. This occurred due to the electrostatic attraction, consequently impacting the hydration characteristics of the quartz surface.



INTRODUCTION Water molecules are adsorbed on the hydrophilic mineral’s surface (e.g., quartz and clay minerals) and arranged in a certain order which is referred to as mineral hydration. This process plays an important role in colloidal chemistry,1 wastewater treatment,2 and mineral processing.3 For instance, with highly muddied coal slurry containing a large amount of hydrophilic quartz and clay mineral particles, slurry’s sedimentation and dewatering becomes more difficult owing to mineral hydration.4,5 The reason for this is the strong hydration repulsion and stereohindrance effect being produced and making the dispersion system stable.6−9 For this reason, the mineral surface hydration characteristics have emerged as the critical aspect of solving problems dealing with how to treat coal slurry. Mineral hydration in aqueous solution has been extensively studied using SFA,10,11 X-ray diffraction,12,13 transmission electron microscopy,14 NMR,15,16 atomic force microscopy (AFM),17 and other experimental methods.18−22 Referring to the hydration model, Manciu et al.23,24 proposed a polarization strategy, wherein an “ice-like” hydration layer structure appeared on the quartz surface in aqueous solution. The mineral’s hydration layer is characterized by its elasticity and strength, and the water molecules in this region have larger density and viscosity when compared to bulk water.25−28 The thickness of the hydration layer on the silica surface was measured using two methods, namely, viscosity29 and AFM.30 © 2019 American Chemical Society

It was reported that the hydration layer thickness of the clay mineral varied in the ∼0−50 nm range. Density functional theory (DFT) is an effective theoretical tool for calculating the mineral structure and its surface adsorption,31−33 and the first-principles calculation method based on DFT can analyze the interaction mechanism of the mineral surfactant34,35 and mineral water36,37 at the molecular/ atomic level. For example, Liu et al.38 reported that the hydration layer of a mineral can prevent direct adsorption of ether amine, dodecylamine ethoxylate (AC1201) ether and DDA on the mineral surface. Peng et al.39 have employed the periodic DFT method to document the adsorption mechanism of water on sodium-montmorillonite (001) basal and (010) edge surfaces. The atomic structure, preferred adsorption sites, adsorption energies, and vibrational frequencies for water adsorption on the α-quartz (101) surface have been investigated by Bandura et al.40 They noted that the structure of water adsorption on the α-quartz (101) surface was defined by the comprehensive effect of the interactions involving water−water, water−silanol, and silanol-silanol hydrogen bonds. The hydroxylated surfaces were stabilized by weak hydrogen bonds that were evident between the silanol groups and surface-bridging oxygens. Received: May 28, 2019 Accepted: July 12, 2019 Published: July 26, 2019 12711

DOI: 10.1021/acsomega.9b01570 ACS Omega 2019, 4, 12711−12718

ACS Omega

Article

Generally, metal ions are commonly found in flotation pulp and coal slurry. With that, the activation mechanism of quartz by Ca(II) was revealed in the atomic level using the DFT method. It was reported that the Ca(OH)+ ions can repulse the hydration shell and consequently adsorb on the quartz surface; the Ca(OH)+ ion acts as a bridge between the αquartz (101) surface and α-Bromolauric acid (CH3(CH2)9CHBrCOOH, a-BLA).41 Despite the fact that water molecules’ adsorption on the minerals’ surface is generally understood, the effects of various metallic ions (e.g., Na+, Mg2+, Ca2+, etc.) on the quartz surface remain largely unexplored. This requires further investigation and consequently, in this paper, we have undertaken a systematic theoretical study on the adsorption characteristics of a single water molecule on the α-quartz (001) surface with and without the presence of Na+, Mg2+, and Ca2+. It has been done employing the periodic DFT method. The frontier orbital analysis, adsorption configurations, adsorption energies, Mulliken atomic charges, the partial densities of states (PDOSs), and the electron transfer are presented and discussed in this research study.

Figure 1. Frontier orbital analysis of the water molecule (a) and αquartz (001) surface (b). The isosurface value are 0.02 and 0.05 electrons/Å3, respectively.

Considering the periodicity and symmetry of the optimized α-quartz (001) surface, the adsorption sites of the individual water molecule on the quartz surface were predicted. One hollow (H), five top (T), and eight bridge (B) adsorption sites (the dotted line and real line are symmetrical) were considered, according to the number of possible hydrogen bonds as shown in Figure 2.



RESULTS AND DISCUSSION Frontier Orbital Analysis. The frontier orbital theory42 asserts that the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) are the most reactive. In Table 1 mentioned below, the orbital energy Table 1. Frontier Orbital Energy and HOMO−LUMO Gap (ΔE) between the Water Molecule and α-Quartz (001) Surface frontier orbital energy/a.u H2O α-quartz (001)

HOMO−LUMO gap/a.u

HOMO

LUMO

ΔE1

−0.238205 −0.28325

−0.010765 −0.07246

0.16575

ΔE2 0.27248

and HOMO−LUMO gap (ΔE, calculated by eq 1) of the reactants are summarized. l H 2O α ‐ quartz(001) o | o o ΔE1 = |E HOMO − E LUMO m o o H 2O α ‐ quartz(001) o ΔE2 = |E HOMO − E LUMO | n H 2O E HOMO

(1) Figure 2. Initial adsorption sites of the water molecule on the αquartz (001) surface.

H 2O E LUMO

where and represent the orbital energy of the water molecule. Similarly, Eα‑quartz(001) and Eα‑quartz(001) are the HOMO LUMO orbital energy of the α-quartz surface. According to the frontier orbital theory, the smaller the ΔE is, the stronger is the interaction. ΔE1 < ΔE2 indicates that the electrons of the HOMO orbital of the water molecule were transferred to the empty LUMO orbital of the α-quartz (001) surface. It can therefore be predicted that the adsorption of the water molecule on the quartz surface is due to the interaction between the hydrogen atoms of the water molecule and the oxygen atoms of the α-quartz (001) surface. The frontier orbital properties of the water molecule and the α-quartz (001) surface are shown in Figure 1a,b, respectively. It is evident that the HOMO orbital of the water molecule appears on the oxygen atom, while the LUMO orbital of the α-quartz (001) appears on both hydrogen and oxygen atoms located on the surface.

Adsorption Energies and Configuration. Using the following formula below, the adsorption energies (Eads) of the water molecule on the α-quartz (001) surface were calculated Eads = E T − (EQ − E W )

(2)

where ET is the total energy of the water and the α-quartz (001) surface. EW is the energy of an isolated water molecule. EQ is the energy of the α-quartz (001) surface. The adsorption energies (Eads) of the water molecule on the α-quartz (001) surface at different initial positions were calculated using eq 2 and plotted as a function of different positions as illustrated in Figure 3. It can be seen that the Eads of the water molecule on the α-quartz (001) surface range from −16.46 to −72.60 kJ/mol; a similar amount of adsorption 12712

DOI: 10.1021/acsomega.9b01570 ACS Omega 2019, 4, 12711−12718

ACS Omega

Article

Figure 3. Adsorption energies of the single water molecule on the αquartz (001) surface at different initial positions.

Figure 5. Electron density difference of the optimal equilibrium configurations of single water adsorbed on the α-quartz (001) surface; (a) top view; (b) side view. The isosurface value is 0.02 electrons/Å3.

energy (−75.8 kJ/mol) was documented by de Leeuw et al.43 The strong interaction between the water molecule and the quartz surface is indicated here. The optimal adsorption configuration of the single water molecule on the quartz surface is illustrated in Figure 4. In this

The Mulliken charge analysis was conducted on the atoms directly involved in the adsorption of the water molecule on the α-quartz (001) surface, and results are summarized in Table 2. From Table 2, it can be observed that the 1s state of Table 2. Mulliken Charge Populations of Interacted Atoms before and after Adsorption Mulliken charge atomic location H2O

atomic number

adsorption states

s

p

charge/e

Hw

Before After before after before after before after

0.48 0.51 1.89 1.86 0.48 0.52 1.85 1.84

0 0 5.16 5.16 0 0 5.24 5.23

0.52 0.49 −1.05 −1.01 0.52 0.48 −1.09 −1.07

Ow α-quartz (001)

Hs Os

Figure 4. Spatial position diagram of the optimal equilibrium configurations of the single water molecule on the α-quartz (001) surface; (a) top view; (b) side view.

the hydrogen atom (Hw) of the water molecule lost 0.03e while interacting with the oxygen atoms (Os) on the α-quartz (001) surface. In contrast, the hydrogen atom (Hs) of the αquartz (001) surface lost 0.04e during the interaction with oxygen atoms (Ow) of the water molecule. Meanwhile, the oxygen atom (Os) of the α-quartz (001) surface and the oxygen atom (Ow) of the water molecule obtained the corresponding electrons. Based on this observation, it can be concluded that the hydrogen bond is responsible for adsorbing the water molecule on the α-quartz (001) surface. Partial Density of States Analysis. Figures 6 and 7 illustrate the PDOS of the interaction atoms between the water molecule and the α-quartz (001) surface in the optimal equilibrium configuration. The zero point denoted Fermi level EF. It can be seen in Figure 6 that the energy of Os 2p has a negative shifting tendency during the water molecule adsorption process. Bonding effects were evident in the region of −10 ≤ energy/eV ≤ −5 (Figure 6) and −8 ≤ energy/eV ≤ −5 (Figure 7) for Hw 1s and Os 2p, and Hs 1s and OW 2p, respectively. Hence, the existence of the hydrogen bond (Ow− Hs) and (Os−Hw) between the water molecule and the −OH group of the α-quartz (001) surface was observed (see the insets of Figures 6 and 7). The peak intensity ratio of Hw 1s (Figure 6) was found to be larger when compared to the Hs 1s at −6.4 eV, which indicates that the Os−Hw hydrogen bond was stronger.

present study, the interacted atoms of the water molecule and quartz surface were Hw, Ow, and Hs, Os, respectively. It is observed that the water molecule was vertically adsorbed on the α-quartz (001) surface. The optimal adsorption site (i.e., the hydrogen atom, Hw and oxygen atom, Ow of the water molecule interacting with the adjacent oxygen atom, Os and hydrogen atom, and Hs on the nearby quartz surface, respectively) of the water molecule on the α-quartz (001) surface was found at initial position B 6. Charge Analysis. The electron density difference and Mulliken charge populations were investigated to discover the adsorption mechanism of the single water molecule on the αquartz (001) surface. Figure 5 depicts the electron density difference of the adsorption system of the water molecule on the α-quartz (001) surface. In the plot of electron density difference, the blue and yellow regions between two atoms represented the accumulation and depletion of electrons, respectively. It appears that a significant charge was transferred between the water molecule and adjacent atoms of the αquartz (001) surface. The electrons were transferred from the hydrogen atoms of the water molecule to the α-quartz (001) surface. 12713

DOI: 10.1021/acsomega.9b01570 ACS Omega 2019, 4, 12711−12718

ACS Omega

Article

Figure 6. PDOS of the interacted atoms (Hw−Os) between water and the α-quartz (001) surface in the optimum adsorption configuration.

Figure 7. PDOS of the interacted atoms (Hs−Ow) between water and the α-quartz (001) surface in the optimum adsorption configuration.

Figure 8. Initial adsorption position (a), adsorption equilibrium position (b), and electron density difference (c) of water molecule adsorption on the α-quartz (001) surface in the presence of Na+, Mg2+, and Ca2+ (the isosurface value is 0.02 electrons/Å3). 12714

DOI: 10.1021/acsomega.9b01570 ACS Omega 2019, 4, 12711−12718

ACS Omega

Article

Table 3. Mulliken Charge Population of the Atoms in the Presence of Metal Ions

adsorption system H2O

Eads (kJ/mol)

Ca2+

Mulliken charge population

Mulliken charge population

Mulliken charge population

adsorption state

s

p

charge/e

s

p

charge/e

s

Hw

before after before after before after before after before after

0.46 0.56 1.86 1.86 0.52 0.56 1.84 1.83 3.00 2.60 −172.56

0 0 5.16 5.10 0 0 5.23 5.22 6.00 6.00 −131.32

0.54 0.44 −1.01 −0.96 0.48 0.44 −1.07 −1.05 0 0.40 −173.36

0.46 0.55 1.86 1.87 0.52 0.53 1.84 1.83 2.00 1.72

0 0 5.16 5.11 0

0.54 0.45 −1.01 −0.98 0.48 0.47 −1.07 −1.05 0 0.20

0.46 0.53 1.86 1.86 0.52 0.52 1.84 1.84 4.00 3.50

Hs Os

M+/M2+

Mg2+

atomic number

Ow α-quartz (001)

Na+

M

Effect of Na+, Mg2+, and Ca2+ on the Charge Transfer and Adsorption Configurations. Based on the optimal adsorption configuration of the individual water molecule on the α-quartz (001) surface, metal ions were added to the system. The objective here was to investigate the effects of metal ions on the hydration of quartz. It is noted that the simulation parameters were chosen based on the results of water molecule adsorption on the quartz surface. The electron density difference of the water molecule and α-quartz (001) adsorption systems in the presence of different metal ions, namely, Na+, Mg2+, and Ca2+ are shown in Figure 8. It can be seen in Figure 8 that the charge transfers and spatial position between the interacted atoms of the water molecule and quartz surface changed in the presence of metal ions. Although no significant electron accumulation or depletion between the Na+ and the adjacent atoms occurred, what did occur was the phenomenon of electron transfer between Mg2+, Ca2+, and oxygen atom of the water molecule. This happened due to the electrostatic interactions. It is worth noting that the distance (1.598 Å) between Ow and Hs (shown in Figure 4) has been significantly increased (∼8.39 to 77.78%) in the presence of metal ions, namely, 1.732 Å (Mg2+), 1.734 Å (Na+), and 2.841 Å (Ca2+). In contrast, the distance (1.688 Å) between Hw and Os slightly decreased in the presence of metal ions, specifically 1.629 Å (Na+), 1.674 Å (Mg2+). This changed to 2.504 Å when Ca2+ was present. The spatial position of the water molecule appeared to be deflected. The adsorption energies in the presence of Na+, Mg2+ and Ca2+ became, respectively, −172.56, −131.32, and −173.36 kJ/mol. From this it can be stated conclusively that in the presence of Na+ and Ca2+, the adsorption system become more stable, and furthermore the interaction of the water molecule with the surface proved to be stronger. This in turn promoted hydration of the quartz surface and demonstrated that the influence of Mg2+ is weaker than that of Na+ and Ca2+. In the presence of metal ions, the Mulliken charge population of the atoms in the adsorption system is shown in Table 3. The analysis shows that the Hs 1s state interacts with the Ow and lost (0.00−0.04) electron, while the Hw 1s state lost (0.07−0.10) electrons when metal ions were present. In contrast, the Mulliken charge population of the Os revealed no significant changes before and after the adsorption of the water molecule. The water molecule obtained a few electrons, and the Ow 2p state gained (0.01−0.05) electrons. In the presence of metal ions, the charge transfers (0.04−0.07e)

5.23 5.22 6.00 6.08

p

d

charge/e

0 0.23

0.54 0.47 −1.01 −1.00 0.48 0.48 −1.07 −1.05 0 0.27

5.16 5.14

5.23 5.21 6.00 6.00

between interacted atoms in the adsorption system were found to be larger than that (0.00−0.40e) without metal ions. The electrons were transferred from the water molecule to the αquartz (001) surface, causing the quartz surface to become negatively charged. The positively charged metal ions and negatively charged quartz surfaces were adsorbed and this was explained by the electrostatic attraction. The adsorption of metal ions on the surface of the quartz enhanced the interaction between the water molecule and the surface, while the adsorbed metal ions connected the water molecule and surface. Subsequently, these actions promoted hydration of the quartz surface.



CONCLUSIONS This study investigated the effects of metal ions, specifically Na+, Mg2+, and Ca2+ on the adsorption of the single water molecule on the α-quartz (001) surface. To do this, the DFT method was employed. The main findings of the present study are summarized below: (1) The DFT calculation results show that the adsorption energies of the single water molecule on different initial positions of the α-quartz (001) surface ranged from −72.60 to −16.46 kJ/mol, and the corresponding adsorption energy of the optimal adsorption configuration was −72.60 kJ/mol. The water molecule spontaneously formed two hydrogen bonds (i.e., Hw−Os and Hs−Ow) with the surface to adsorb on the α-quartz (001) surface. (2) In the presence of metal ions, the spatial position of the water molecule is deflected, revealing a tendency to move away from Hs and closer to Ow. The Ow−Hs lengths change from 1.598 to 1.734, 1.732, and 2.841 Å in the presence of Na+, Mg2+, and Ca2+, respectively. Meanwhile, the Hw−Os lengths (1.688 Å) change to 1.629, 1.674, and 2.504 Å. The main hydrogen bonding effect is Hw−Os. It appeared that the adsorption systems became stronger in the presence of Na+ and Ca2+, while a relatively weak promotion comes from Mg2+. (3) The surface hydration mechanism of the α-quartz (001) surface occurs mainly through the adjacent water molecules which are adsorbed on the quartz surface by forming hydrogen bonds. Metal ions interact with the quartz surface via electrostatic interaction, subsequently influencing the adsorption of the water molecule and surface hydration characteristics. 12715

DOI: 10.1021/acsomega.9b01570 ACS Omega 2019, 4, 12711−12718

ACS Omega

Article

Table 4. Comparison between the Experimental and Computational Lattice Parameters of Optimized Bulk α-Quartz lattice parameters



computational parameters

a/Å

b/Å

c/Å

α/(deg)

β/(deg)

γ/(deg)

references

GGA/700 eV GGA/1000 eV experimental values GGA/400 eV

5.05 5.09 4.92 4.93

5.05 5.09 4.92 4.93

5.54 5.58 5.41 5.44

90 90 90 90

90 90 90 90

120 120 120 120

44 47 46 this work

study were H 1s1, O 2s22p4, Na 2s22p63s1, Mg 2s22p63s2, and Ca 3s23p64s2. The Broyden−Fletcher−Goldfarb−Shanno (BFGS) algorithm52 helped to optimize the atomic positions. The Tkatchenko−Scheffler van der Waals correction method53 was employed to correct density functional calculations for the missing van der Waals interactions. With computation accuracy and efficiency being taken into account, the kinetic energy cut-off for the plane wave basis was set to 400 eV, and the symmetrized Brillouin zone was obtained with a 2 × 2 × 1 Monkhorst−Pack grid according to a series of convergence tests (see Figure S1 in the Supporting Information). Regarding other calculations, the tolerances of energy, force, stress, and displacement were 1.0 × 10−5 eV/atom, 0.05 eV/Å, 0.05 GPa, and 0.001 Å, respectively. Finally, the Dmol3 module was employed to calculate the frontier orbital of water and quartz. In other words, after optimization in CASTEP, the frontier orbital energy could be obtained by calculating the optimized unit cell with singlepoint energy and Gamma point (K-point) in DMol3. The parameters selected were shown as follows: exchange correlation function GGA-PBE, effective core potential and DNP basis set, fine precision, and convergence standard of the self-consistent field 1.0 × 10−6 eV/atom.

COMPUTATIONAL METHODS Models. In quartz, the silica tetrahedra are bonded in the three-dimensional framework. Each tetrahedron is bonded to four other tetrahedra (with an oxygen atom shared at every corner of each tetrahedron). The geometric optimization of the quartz primitive cell was conducted and the results are summarized in Table 4. Compared with the previous simulations44,45 and experimental46 results, the α-quartz lattice parameters obtained in this simulation are comparable to the experimental values. Unlike many brittle crystals, α-quartz does not possess perfect crystallographic planes of cleavage, and the (001) surface emerges as the most stable surface as predicted by both interatomic potential48 and DFT methods.43 While the fresh αquartz (001) surface is easily hydroxylated in aqueous solution, the top and side views of the hydroxylated α-quartz (001) surface are shown in Figure 9a,b, respectively. The periodic supercell (2 × 2 × 1) was used with a vacuum thickness of 10 Å to avoid the interaction between the adjacent levels.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.9b01570. Batch convergence tests of bulk α-quartz to obtain the reasonable cut-off energy and K-point (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: ff[email protected]. ORCID

Chunfu Liu: 0000-0003-1506-4712 Fanfei Min: 0000-0001-9161-745X Lingyun Liu: 0000-0002-8676-097X

Figure 9. Top (a) and side (b) views of the hydroxylated α-quartz (001) surface.

Notes

The authors declare no competing financial interest.



Computational Details. DFT calculations were implemented in the CASTEP program (Materials Studio version 8.0 software, Accerly Corporation).49 Electron exchange and correlation were described using the Perdew−Burke−Ernzerhof (PBE) function of the generalized gradient approximation (GGA).50 The interaction between valence electrons and the ionic core was described by the ultrasoft pseudopotential,51 and the valence electron configurations considered in this

ACKNOWLEDGMENTS The financial support for this work was provided by the National Natural Science Foundation of China under grant no. 51874011 and the National Natural Science Foundation of China under grant no. 51804009. Both sources are gratefully acknowledged for their contribution to this study. 12716

DOI: 10.1021/acsomega.9b01570 ACS Omega 2019, 4, 12711−12718

ACS Omega



Article

mineral−water interfaces: Application to cement materials. Cem. Concr. Res. 2007, 37, 337−347. (22) Delville, A. Structure and Properties of Confined Liquids - a Molecular-Model of the Clay Water Interface. J. Phys. Chem. 1993, 97, 9703−9712. (23) Manciu, M.; Ruckenstein, E. The polarization model for hydration/double layer interactions: the role of the electrolyte ions. Adv. Colloid Interface Sci. 2004, 112, 109−128. (24) Manciu, M.; Calvo, O.; Ruckenstein, E. Polarization model for poorly-organized interfacial water: hydration forces between silica surfaces. Adv. Colloid Interface Sci. 2006, 127, 29−42. (25) Perry, T. D.; Cygan, R. T.; Mitchell, R. Molecular models of a hydrated calcite mineral surface. Geochim. Cosmochim. Acta 2007, 71, 5876−5887. (26) Fenter, P.; Lee, S. S. Hydration layer structure at solid−water interfaces. MRS Bull. 2014, 39, 1056−1061. (27) Martin-Jimenez, D.; Chacon, E.; Tarazona, P.; Garcia, R. Atomically resolved three-dimensional structures of electrolyte aqueous solutions near a solid surface. Nat. Commun. 2016, 7, 12164. (28) Sun, C. Q.; Sun, Y. Superlubricity of Ice. The Attribute of Water; Springer, 2016; pp 203−243. (29) Song, S.; Peng, C. Thickness of Solvation Layers on Nano-scale Silica Dispersed in Water and Ethanol. J. Dispersion Sci. Technol. 2005, 26 (2), 197−201. (30) Peng, C.; Song, S. Determination of thickness of hydration layers on mica in aqueous solutions by using AFM. Surf. Rev. Lett. 2004, 11, 485−489. (31) Chen, J.; Min, F.-f.; Liu, L.; Liu, C.; Lu, F. Experimental investigation and DFT calculation of different amine/ammonium salts adsorption on kaolinite. Appl. Surf. Sci. 2017, 419, 241−251. (32) McKenzie, M. E.; Goyal, S.; Lee, S. H.; Park, H.-H.; Savoy, E.; Rammohan, A. R.; Mauro, J. C.; Kim, H.; Min, K.; Cho, E. Adhesion of Organic Molecules on Silica Surfaces: A Density Functional Theory Study. J. Phys. Chem. C 2016, 121, 392−401. (33) Zhao, C.; Chen, J.; Li, Y.; Huang, D. W.; Li, W. DFT study of interactions between calcium hydroxyl ions and pyrite, marcasite, pyrrhotite surfaces. Appl. Surf. Sci. 2015, 355, 577−581. (34) Xie, J.; Li, X.; Mao, S.; Li, L.; Ke, B.; Zhang, Q. Effects of structure of fatty acid collectors on the adsorption of fluorapatite (0 0 1) surface: A first-principles calculations. Appl. Surf. Sci. 2018, 444, 699−709. (35) Gattinoni, C.; Ewen, J. P.; Dini, D. Adsorption of Surfactants on α-Fe2O3(0001): A Density Functional Theory Study. J. Phys. Chem. C 2018, 122, 20817−20826. (36) Chen, J.; Min, F.-f.; Liu, L.-y.; Liu, C.-f. Mechanism research on surface hydration of kaolinite, insights from DFT and MD simulations. Appl. Surf. Sci. 2019, 476, 6−15. (37) Zhang, L.; Wu, Y.; Liu, Y.; Li, H. DFT study of single water molecule adsorption on the (100) and (101) surfaces of KH2PO4. RSC Adv. 2017, 7, 26170−26178. (38) Liu, A.; Fan, J.-c.; Fan, M.-q. Quantum chemical calculations and molecular dynamics simulations of amine collector adsorption on quartz (0 0 1) surface in the aqueous solution. Int. J. Miner. Process. 2015, 134, 1−10. (39) Peng, C.; Min, F.; Liu, L.; Chen, J. A periodic DFT study of adsorption of water on sodium-montmorillonite (001) basal and (010) edge surface. Appl. Surf. Sci. 2016, 387, 308−316. (40) Bandura, A. V.; Kubicki, J. D.; Sofo, J. O. Periodic Density Functional Theory Study of Water Adsorption on the alpha-Quartz (101) Surface. J. Phys. Chem. C 2011, 115, 5756−5766. (41) Zhu, Y.; Luo, B.; Sun, C.; Liu, J.; Sun, H.; Li, Y.; Han, Y. Density functional theory study of α-Bromolauric acid adsorption on the α-quartz (1 0 1) surface. Miner. Eng. 2016, 92, 72−77. (42) Yoshizawa, K.; Tada, T.; Staykov, A. Orbital views of the electron transport in molecular devices. J. Am. Chem. Soc. 2008, 130, 9406−9413. (43) de Leeuw, N. H.; Higgins, F. M.; Parker, S. C. Modeling the surface structure and stability of α-quartz. J. Phys. Chem. B 1999, 103, 1270−1277.

REFERENCES

(1) Israelachvili, J.; Wennerström, H. Role of hydration and water structure in biological and colloidal interactions. Nature 1996, 379, 219−225. (2) Espantaleón, A.; Nieto, J. A.; Fernandez, M.; Marsal, A. Use of activated clays in the removal of dyes and surfactants from tannery waste waters. Appl. Clay Sci. 2003, 24, 105−110. (3) Miller, J. D.; Wang, X.; Jin, J.; Shrimali, K. Interfacial water structure and the wetting of mineral surfaces. Int. J. Miner. Process. 2016, 156, 62−68. (4) Liu, C.; Min, F.; Liu, L.; Chen, J.; Du, J. Mechanism of hydrolyzable metal ions effect on the zeta potential of fine quartz particles. J. Dispersion Sci. Technol. 2018, 39, 298−304. (5) Peng, C.; Zhong, Y.; Min, F. Adsorption of alkylamine cations on montmorillonite (001) surface: A density functional theory study. Appl. Clay Sci. 2018, 152, 249−258. (6) Israelachvili, J. N.; McGuiggan, P. M. Forces Between Surfaces in Liquids. Science 1988, 241, 795−800. (7) Israelachvili, J. N. Intermolecular and Surface Forces; Academic Press, 2011. (8) Peng, C.; Song, S.; Fort, T. Study of hydration layers near a hydrophilic surface in water through AFM imaging. Surf. Interface Anal. 2006, 38, 975−980. (9) Min, F.; Peng, C.; Liu, L. Investigation on hydration layers of fine clay mineral particles in different electrolyte aqueous solutions. Powder Technol. 2015, 283, 368−372. (10) Pashley, R. M. DLVO and Hydration Forces between Mica Surfaces in Li+, Na+, K+, and Cs+ Electrolyte-Solutions - a Correlation of Double-Layer and Hydration Forces with Surface Cation-Exchange Properties. J. Colloid Interface Sci. 1981, 83, 531−546. (11) Pashley, R. M.; Israelachvili, J. N. DLVO and hydration forces between mica surfaces in Mg2+, Ca2+, Sr2+, and Ba2+ chloride solutions. J. Colloid Interface Sci. 1984, 97, 446−455. (12) Ferrage, E.; Lanson, B.; Michot, L. J.; Robert, J.-L. Hydration properties and interlayer organization of water and ions in synthetic Na-smectite with tetrahedral layer charge. Part 1. Results from X-ray diffraction profile modeling. J. Phys. Chem. C 2010, 114, 4515−4526. (13) Ferrage, E.; Lanson, B.; Malikova, N.; Plançon, A.; Sakharov, B. A.; Drits, V. A. New Insights on the Distribution of Interlayer Water in Bi-Hydrated Smectite from X-ray Diffraction Profile Modeling of 00l Reflections. Chem. Mater. 2005, 17, 3499−3512. (14) Tadjiev, D. R.; Hand, R. J.; Zeng, P. Comparison of glass hydration layer thickness measured by transmission electron microscopy and nanoindentation. Mater. Lett. 2010, 64, 1041−1044. (15) Faux, D. A.; Cachia, S. H. P.; McDonald, P. J.; Bhatt, J. S.; Howlett, N. C.; Churakov, S. V. Model for the interpretation of nuclear magnetic resonance relaxometry of hydrated porous silicate materials. Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys. 2015, 91, 032311. (16) Grangeon, S.; Claret, F.; Roosz, C.; Sato, T.; Gaboreau, S.; Linard, Y. Structure of nanocrystalline calcium silicate hydrates: insights from X-ray diffraction, synchrotron X-ray absorption and nuclear magnetic resonance. J. Appl. Crystallogr. 2016, 49, 771−783. (17) Miyazawa, K.; Kobayashi, N.; Watkins, M.; Shluger, A. L.; Amano, K.-i.; Fukuma, T. A relationship between three-dimensional surface hydration structures and force distribution measured by atomic force microscopy. Nanoscale 2016, 8, 7334−7342. (18) Flörsheimer, M.; Kruse, K.; Polly, R.; Abdelmonem, A.; Schimmelpfennig, B.; Klenze, R.; Fanghänel, T. Hydration of mineral surfaces probed at the molecular level. Langmuir 2008, 24, 13434− 13439. (19) Salles, F.; Douillard, J.-M.; Denoyel, R.; Bildstein, O.; Jullien, M.; Beurroies, I.; Van Damme, H. Hydration sequence of swelling clays: evolutions of specific surface area and hydration energy. J. Colloid Interface Sci. 2009, 333, 510−522. (20) Leng, Y.; Cummings, P. T. Hydration structure of water confined between mica surfaces. J. Chem. Phys. 2006, 124, 074711. (21) Kalinichev, A. G.; Wang, J.; Kirkpatrick, R. J. Molecular dynamics modeling of the structure, dynamics and energetics of 12717

DOI: 10.1021/acsomega.9b01570 ACS Omega 2019, 4, 12711−12718

ACS Omega

Article

(44) Goumans, T. P. M.; Wander, A.; Brown, W. A.; Catlow, C. R. A. Structure and stability of the (001) alpha-quartz surface. Phys. Chem. Chem. Phys. 2007, 9, 2146−2152. (45) Plessow, P. N.; Sánchez-Carrera, R. S.; Li, L.; Rieger, M.; Sauer, S.; Schaefer, A.; Abild-Pedersen, F. Modeling the Interface of Platinum and alpha-Quartz(001): Implications for Sintering. J. Phys. Chem. C 2016, 120, 10340−10350. (46) Levien, L.; Prewitt, C. T.; Weidner, D. J. Structure and Elastic Properties of Quartz at Pressure. Am. Mineral. 1980, 65, 920−930. (47) Plessow, P. N.; Sánchez-Carrera, R. S.; Li, L.; Rieger, M.; Sauer, S.; Schaefer, A.; Abild-Pedersen, F. Modeling the Interface of Platinum and α-Quartz (001): Implications for Sintering. J. Phys. Chem. C 2016, 120, 10340−10350. (48) Rignanese, G.-M.; De Vita, A.; Charlier, J.-C.; Gonze, X.; Car, R. First-principles molecular-dynamics study of the (0001) alphaquartz surface. Phys. Rev. B 2000, 61, 13250−13255. (49) Segall, M. D.; Lindan, P. J. D.; Probert, M. J.; Pickard, C. J.; Hasnip, P. J.; Clark, S. J.; Payne, M. C. First-principles simulation: ideas, illustrations and the CASTEP code. J. Phys.: Condens. Matter 2002, 14, 2717−2744. (50) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865−3868. (51) Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B: Condens. Matter Mater. Phys. 1990, 41, 7892−7895. (52) Liu, D. C.; Nocedal, J. On the Limited Memory Bfgs Method for Large-Scale Optimization. Math. Program. 1989, 45, 503−528. (53) Bučko, T.; Lebegue, S.; Hafner, J.; Angyan, J. G. TkatchenkoScheffler van der Waals correction method with and without selfconsistent screening applied to solids. Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 87, 064110.

12718

DOI: 10.1021/acsomega.9b01570 ACS Omega 2019, 4, 12711−12718