Water Adsorption on Clean and Defective Anatase TiO2 (001

Mar 29, 2018 - We use ab initio molecular dynamics simulations to study the adsorption of thin water films with 1 and 2 ML coverage on anatase TiO2 (0...
3 downloads 0 Views 3MB Size
Article pubs.acs.org/JPCB

Cite This: J. Phys. Chem. B XXXX, XXX, XXX−XXX

Water Adsorption on Clean and Defective Anatase TiO2 (001) Nanotube Surfaces: A Surface Science Approach Stephane Kenmoe,*,† Oleg Lisovski,‡ Sergei Piskunov,‡ Dmitry Bocharov,‡ Yuri F. Zhukovskii,‡ and Eckhard Spohr†,§ †

Department of Theoretical Chemistry, University of Duisburg-Essen, Universitätsstr. 2, D-45141 Essen, Germany Institute of Solid State Physics, University of Latvia, 8 Kengaraga Str., Riga LV-1063, Latvia § Center of Computational Sciences and Simulation, University of Duisburg-Essen, D-45141 Essen, Germany ‡

S Supporting Information *

ABSTRACT: We use ab initio molecular dynamics simulations to study the adsorption of thin water films with 1 and 2 ML coverage on anatase TiO2 (001) nanotubes. The nanotubes are modeled as 2D slabs, which consist of partially constrained and partially relaxed structural motifs from nanotubes. The effect of anion doping on the adsorption is investigated by substituting O atoms with N and S impurities on the nanotube slab surface. Due to strain-induced curvature effects, water adsorbs molecularly on defect-free surfaces via weak bonds on Ti sites and H bonds to surface oxygens. While the introduction of an S atom weakens the interaction of the surface with water, which adsorbs molecularly, the presence of an N impurity renders the surface more reactive to water, with a proton transfer from the water film and the formation of an NH group at the N site. At 2 ML coverage, a further surface-assisted proton transfer takes place in the water film, resulting in the formation of an OH− group and an NH+2 cationic site on the surface.



INTRODUCTION Photocatalytic water splitting is considered as a clean and environmentally responsible way of satisfying global energy demands. For this purpose, the ability of different photocatalysts to drive the water splitting reaction under sunlight illumination has been studied. Among these, TiO2 is one of the most-studied materials due to many beneficial properties, such as chemical and optical stability, nontoxicity, low cost, and abundance.1 The higher efficiency of anatase among TiO2 polymorphs in photocatalytic applications2 and the presence of water in operando conditions have motivated a considerable amount of studies on the anatase TiO2/water interface for a better understanding of the elementary processes and for the improvement of its photocatalytic properties.3 Crystalline anatase TiO2 grows naturally as a truncated octahedron exposing mainly the (101) and (001) facets, with the (101) facet covering more than 90% of the surface.4,5 The interaction of water with the predominant (101) and the less exposed (001) facets have been addressed both theoretically and experimentally.3,6,7 The studies agree on the fact that the (001) facet shows a chemical reactivity higher than that of the (101) facet. The reason is that, unlike the (101) facet, all of the surface species are undercoordinated on the (001) facet, conferring it a higher surface energy, as shown in the Wulff construction for anatase TiO2.5 While on the (101) facet molecular adsorption at the 5-fold coordinated Ti sites is favored from submonolayer coverage to © XXXX American Chemical Society

bulk water, dissociative adsorption is prominent on the (001) facet.8,9 Using a (2 × 2) supercell, Vittadini et al.8 showed that water fully dissociates on the (001) surface at submonolayer coverage, while at monolayer coverage, only one-half of the water molecules adsorb dissociatively at the 5-fold-coordinated Ti sites (Ti5c), with the OH radicals forming H bonds with the remaining water molecules located in an upper layer. A later study by Sumita et al.9 showed that the adsorption mode in the contact water layer depends on the size of the supercell, as two consecutive Ti5c sites in the direction of [100] are required for dissociative adsorption to occur. Using a (3 × 3) supercell to model the interface of the TiO2 (001) surface with bulk water in this latter study, they found that intact water molecules and OH groups are present in the contact layer, forming strong and weak H bonds to molecules in a relatively mobile upper layer. In accordance with previous studies on water adsorption on TiO2 surfaces, this led to the conclusion that at low temperatures, the second or third layer in the water films could already be considered as similar to the free water surface to the vacuum in a thicker film. Though efforts in characterizing the interface between water and TiO2 (001) 2D surfaces have hoisted the understanding of Special Issue: Ken A. Dill Festschrift Received: November 28, 2017 Revised: March 15, 2018 Published: March 29, 2018 A

DOI: 10.1021/acs.jpcb.7b11697 J. Phys. Chem. B XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry B interfacial water as well as the properties of the TiO2 films3 to a creditable level, very little is known about these properties for substrates of lower dimensionality. One of such are three monolayer thick TiO2 (001) nanotubes, which are more stable than the corresponding 2D films, due to negative strain energy.10 Also, nanostructuring TiO2, e.g., by the formation of TiO2 nanotubes (NTs) with large surface areas, can enhance the photocatalytic activity. In this respect, TiO2 (001) nanotubes were shown to be a promising candidate for photocatalytic water splitting, as their visible-light-driven photocatalytic response can be tuned by N and S codoping.11 Band gap engineering using anion impurities is known to be a practical way to tailor the optical response of nanotubes.12 Particularly, N and S substitutional impurities increase the visible-light-driven photocatalytic response of a TiO2 (001) nanotube, suitably aligning its bands relative to the water redox potential.11 In reality, these nanotubes operate in an aqueous environment, hence the importance of simulating their surface in contact with water and their dynamic behavior. Yet information about the water/TiO2 NT interface is scarce, because of the difficulty to characterize this interface both at the theoretical and at the experimental levels. From the modeling point of view, the main issue is the size of the structures, since high photocatalytic performance can be expected, e.g., from a (001) TiO2 (36,0) nanotube codoped with S and N, which then contains 648 atoms in a ring-like nanotube with a diameter of 5 nm11 and a length of 0.7 nm, requiring a concomitant large number of water molecules. High-level quantum chemistry for this model is hardly achievable in aqueous medium, since the computational cost would be very high. Recently, a slab-based surface science approach, in which partially relaxed and partially constrained structural motifs from water-covered cylindrical NTs in a 2D periodic geometry are used, has been proposed as an alternative to circumvent this difficulty.13 Such a scheme, named the CFVS (constrained fixed volume slab) model in this recent work, can reduce the size of the system under study and allow for analysis of dynamics and excitations at the interface, while keeping characteristic nanotube properties. Though the efficiency of this approach is restricted to the description of water adsorption on the outer side of nanotubes with large diameters and relatively thick walls, it is nevertheless suitable for our purpose, since, in reality, most of the surface chemistry and dynamics take place on the outer part of the nanotube. Casarin et al.14 proposed a 1D model in which only a fragment of the nanotube is included, and which enables the study of water interactions with both sides of the nanotube. Such an approach is similar to the curved constrained fixed volume slab (CCFVS) model discussed in our recent study.13 This model retains the original nanotube strain due to curvature and showed a good performance in reproducing the adsorption energy of the original tubular structure on both sides. However, it fails in describing the electronic structure properties, in particular, the densities of states, which exhibit a strong mismatch due to unphysical properties near the kink sites and the ensueing changes due to varying distances from the kink. Experimentally, it is difficult to synthesize large high-quality anatase single crystals with a high percentage of (001) facets. Most of the synthesis methods are time-consuming and expensive. Recently, considerable progress in preparing anatase TiO2 with exposed (001) facets has been achieved by Lu et al.,15 who used hydrofluoric acid as a morphology-controlling agent to synthesize TiO2 single crystals with 47% of (001)

facets. From thereon, many experimental studies have followed, in which single-crystalline anatase TiO2 nanotubes, mainly with the exposed (001) plane, were grown using different hydrothermal routes. Such nanotubes have shown enlarged photocatalytically active surfaces, enhanced conversion efficiencies, and improved electronic transport properties.16,17 Photoinduced superhydrophilicity of anatase TiO 2 nanotube surfaces18−20 have motivated the study of environmental effects on wetting properties. Using contact angle measurements and Fourier transform infrared (FTIR) spectroscopy, Shin and coworkers21 showed that freshly anodized and annealed anatase TiO2 nanotubes have a hydrophilic character. Studying the effect of contaminants upon air exposure, tube diameter, UV illumination, and organic modifications of nanotube surfaces has enabled a better control of factors governing changes in wettability. Although a suitable combination of these factors allow, in principle, tuning the nanotube to acquire a desired wettability, ranging from superhydrophobic to superhydrophilic,18−21 clear insight into the detailed atomistic interactions occurring at the interface is still missing. To shed some light on these open issues, studying the adsorption of water on nanotubes in vacuum is a primordial step toward understanding the complex mechanisms behind the wetting or photocatalytic water splitting processes. In the present article, we address the structural properties, surface chemistry, and dynamics of TiO2 (001) nanotubes covered with thin water films. Using a 2D slab representation of the nanotube, which takes into account the strain and partial curvature of the nanotube, ab initio molecular dynamics calculations are performed to study the distribution of water molecules and the strength of their hydrogen bond network at the interface of the nanotube with water. As the presence of surface OH− groups is crucial in most photocatalytic reactions, the impact of N and S dopants on the degree of dissociation of water molecules is investigated. The effect of water coverage on these properties is investigated by increasing the thickness of the water film up to 2 ML.



COMPUTATIONAL DETAILS The anatase TiO2 (001) nanotube was modeled by a slab of three O−Ti−O units separated by a vacuum region of 18.1 Å. The atoms in the bottom O−Ti−O unit were kept fixed at their initial configurations in the nanotube, while the remaining ones were allowed to relax. The Ti and O atoms in the bottom layer were passivated with pseudo hydrogens. The importance of surface passivation with pseudo hydrogen on the structure and energetics of water-covered thin TiO2 films has previously been demonstrated by Kowalski et al.22 An orthorhombic supercell, with a (3 × 3) periodicity in the lateral directions, x and y, and dimensions of 10.5 Å × 12.6 Å × 25.3 Å, was used to study the interaction of water films with the slab (see Figure S1 in the Supporting Information for a graphical representation of the structure). This 2D slab model is used to simulate the interaction of a (36,0) nanotube of 5 nm diameter with a wall thickness of 6.5 Å, corresponding to the slab thickness, covered by thin water films. It was recently shown that using this slab representation to simulate the adsorption of a water monolayer on the outer side of a TiO2 (001) nanotube with a large radius retains the main structural and energetic features of the nanotube.13 As a reference, we also studied the planar 2D interface. For this, a bigger supercell with dimensions 11.353 Å × 11.353 Å × 28.544 Å was used as in a previous study by Sumita et al.9 Due to the asymmetry imposed by the one-sided B

DOI: 10.1021/acs.jpcb.7b11697 J. Phys. Chem. B XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry B

molecules bind on top of the 5-fold-coordinated Ti surface sites (Ti5c) and form H bonds to the 2-fold-coordinated O surface sites (O2c). On defect-free surfaces, water adsorbs molecularly. The situation is similar on S-doped surfaces, but with the water layer being buckled due to the larger size of the S atom. On Ndoped surfaces, an NH group is present on the N site at 1 ML. At 2 ML, a water molecule in the contact layer dissociates and transfers a proton via the Grothuss mechanism30 to the neighboring NH group. This leads to the formation of an NH+2 cation at the N site and an OH− group at the adjacent Ti5c site. On codoped surfaces, the NH group is also present at the N site. However, no further protonation is observed for this system. On all surfaces, we observe that there is a critical coverage of water molecules. Only 1/3 of the Ti5c atoms interact directly with water molecules. This observation was also made by Sumita et al.,9 where full coverage of Ti5c with a monolayer of water could not be achieved. In the water film, molecules form H bonds to the surface and among each other. At 1 ML, a disordered buckled water layer is formed, increasing the water coverage to 2 ML leads to the formation of a multilayer structure. The density distribution of water molecules along the vertical direction is also investigated. We calculate the number density profiles of water oxygen and hydrogen atoms along the (001) direction, and the positions of Ti5c atoms are taken as reference (Figure 2). All Ti5c density profiles are centered around the 0 Å value at monolayer (1 ML) coverage. At 2 ML water content, all peaks become broader. On the pristine surface and the N-doped surface, the Ti5c density maximum splits into two peaks about 0.4 Å apart. On the N+S-codoped surface, only a shoulder is visible at the larger distance. These features of the Ti5c density are characteristic for the outward relaxation of only some Ti atoms in the presence of water. Note that the pristine surface also shows a small shoulder at about 0.2 Å. The extent of these relaxation phenomena is a consequence of the interaction of the Ti atoms with water oxygen atoms. The O2c density profiles show single peaks localized around 0.6−0.7 Å. Thus, the surface O2c atoms undergo very little relaxation upon adsorption of water on the surface, independent of coverage. This originates from the weaker interaction of O2c sites with water via H bonds relative to the stronger interactions of Ti with water oxygen. Unlike on the two-dimensional flat surface, where the water oxygen profile shows distinct peaks,9 the water oxygen distributions on the nanotube surfaces in Figure 2 are broadened. They consist of double peaks centered at 2.3 and 2.9 Å for 1 ML. The small separation between the two peaks indicates the presence of a buckled water layer at monolayer coverage. In fact, the maximum at 2.9 Å is too far from the surface to be due to direct Ti−water interactions. Rather it appears to be due to water−water interactions. The first peak of the water oxygen density at 2 ML coverage is located almost at the same position as the one at 1 ML. The second peak, however is shifted to about 3.6 Å; it is significantly broadened and extends to large coordinates, which shows the tendency of water molecules to form multilayers. The H atom profile supports the formation of the buckled monolayer feature at 1 ML. It shows a main peak with its maximum between the two oxygen maxima; there is a shoulder on each side of the main hydrogen peak. These shoulders have different relative intensities for the different surfaces. The main hydrogen maximum at about 2.5−2.7 Å is located about 2 Å above the position of the O2c maxima and represents hydrogen

adsorption of water and the different terminations in the slab’s outermost layer and innermost layer, where the atoms are frozen, a dipole correction was introduced to cancel the electric field gradient in the vacuum. The CP2K/Quickstep package23 was used to perform Born− Oppenheimer molecular dynamics simulations and static geometry optimizations. For the latter, the BFGS minimizer was employed. The total energies were calculated at the Γ point only because of the large size of the supercells. The generalized gradient approximation (GGA) within its PBE formulation24 was used to account for exchange correlation effects. It is known that no GGA functional can simultaneously describe the structure, dynamics, and thermodynamics of liquid water.25 Hopes in a fully satisfactory description may be achievable by dispersion-inclusive exchange-correlation functionals.25−27 Therefore, the PBE-D3 dispersion interaction correction as proposed by Grimme28 was used here. Though this functional lacks structural softening due to its overbinding character,25 its efficiency to describe the state of dissociation of water at TiO2 interfaces has been demonstrated.29 The 3s, 3p, 4s, and 3d electrons of Ti atoms and the 2s and 2p electrons of the O atoms were considered as valence electrons, and the Goedecker−Teter−Hutter (GTH) pseudopotentials were used to treat the core electrons. The double-ζ basis functions with one set of polarization functions (DZVP) were used as basis sets with a plane wave cutoff of 400 Ry. A Nosé−Hoover thermostat with a target temperature of 300 K and a time constant of 1 ps was used to impose NVT conditions to the system. A time step of 0.5 fs was used in Born−Oppenheimer molecular dynamics for a total simulation time of 20 ps after equilibration. The impact of dopants on the structure, reactivity, and dynamics of the water films was investigated by replacing surface oxygens by one N, one S, or one N and one S atom simultaneously in the outermost layer of the surface. The N+S configurations were chosen by replacing two surface oxygen atoms 5.6 Å apart along the diagonal direction of the 3 × 3 surface cell. This places the two dopants at the largest possible distance on the surfaces. The procedure corresponds to surface dopant concentrations of 1/18 and 1/9, respectively. To keep the whole system neutral, an extra H atom was added in the water film in case of N or N+S doping.



RESULTS AND DISCUSSIONS We start our investigation by studying the structure of the interface. Figure 1 shows the final snapshots after equilibrium trajectories lasting for 20 ps each. The images are (from left to right) for the interface of water in contact with the clean, the Ndoped, the S-doped, and the N+S-codoped nanotube surfaces at coverages of 1 ML (top) and 2 ML (bottom). Water

Figure 1. Snapshots of equilibrium trajectories for the pristine, Ndoped, S-doped, and N+S-codoped nanotube surfaces covered with a single (top row) or two monolayers of water (bottom row). C

DOI: 10.1021/acs.jpcb.7b11697 J. Phys. Chem. B XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry B

Figure 2. Number density profiles for the defect-free (top left), N-doped (top right), S-doped (bottom left), and N+S-codoped (bottom right) nanotube surfaces. The zero is defined to be located 4.9 Å above the (rigid) bottom Ti layer; it approximately corresponds to the position of the top Ti layer. Lines with symbols are for 1 ML water coverage, lines without symbols are for 2 ML coverage. Note that below about 1.1 Å, the red oxygen curves represent O2c oxygen atoms, above 1.8 Å, they represent water oxygen atoms.

To gain deeper insight into the observed structural properties, characteristic radial distribution functions (RDFs) were calculated. On defect-free surfaces, where molecular adsorption has been observed on Ti5c sites, the peak positions in the Ti5c-Ow RDFs are located at 2.1 Å and 2.2 Å, respectively, at 1 and 2 ML (Figure 3, top). This shows that water binds coordinatively to the Ti5c sites. In the O2c-Hw RDFs (Figure 3, bottom), the first peak is centered around 1.9 Å, which indicates the presence of strong H bonds between the water molecules and the O2c surface sites. Thus, on defect-free surfaces, water binds molecularly via coordinated bonds to the Ti sites and via hydrogen bonds to the oxygen sites. It is known that the state of dissociation of water on TiO2 strongly depends on the supercell size and slab thickness.3 For the anatase (001) surface modeled with a 4 TiO2-layered slab, partial dissociation was found at the interface with bulk water by Sumita et al.9 They found no further change in the structure and adsorption mode upon increasing the slab thickness to 5 layers. Thus, for comparison, we also studied the interface of the 4 TiO2 layered slab surface with a single monolayer of water, using the same supercell model as in ref 9 and the same settings as described in the Computational Details Section. We also observed partial water dissociation in this case (see Figure 4, top left), and the interface looks similar to the contact layer at the interface with bulk water.9 Our nanotube slab model consists of 3 TiO2 layers. Thus, to provide insight into the effect of nanostructuring and curvature on the adsorption, we also simulated the adsorption of a monolayer of water on a 3 TiO2-layered slab surface and still observed partial dissociation (Figure 4, top right). We observe no qualitative change in the surface chemistry of the contact layer when decreasing the thickness from 5 to 3 TiO2 layers or when decreasing the water coverage from a “bulk” film to the monolayer. Thus, we can

atoms that form interwater hydrogen bonds within the buckled (bilayer) of water. The shoulders at around 2.1 Å indicate the formation of strong hydrogen bonds to the O2c surface atoms. The corresponding shoulders around 3.5 Å represent the dangling OH bonds of the water layer pointing away from the surface. At 2 ML, the hydrogen profile is less regular. The atom densities range to about 6 Å and indicate the beginning formation of a multilayer film. Similar features are observed in the case of pure N-doping (Figure 2, top right), except for the presence of an N peak at 0.7−0.9 Å. The N peak is shifted outward by about 0.2 Å, indicating that the N impurity interacts more strongly with the aqueous phase than the regular O2c sites. Furthermore, the small-coordinate tail of the water hydrogen density profile extends almost toward the position of the O2c and N density maxima. This shows that NH and OH bonds are present in the contact layers (see Figure 1). The position of the N peak shifts even further outward from the surface when the water coverage increases to 2 ML. This enhanced interaction is confirmed by the formation of an NH+2 entity. In the presence of the S codopant, no NH+2 is formed, and the N profile remains at positions characteristic for a surface NH group. In fact, S-doping makes the Ti5c sites, in general, less reactive. As can be inferred from Figure 2, S-doping leads only to a single, somewhat broader density maximum, where the characteristic splitting, or the pronounced shoulder, which are present in the sulfur-free case, is absent. Due to less efficient sp hybridization compared to O and N atoms, the S atom density relaxes more strongly (by about 0.8 Å) outward than that of the O2c and N densities. Consequently, the sulfur position at around 1.5 Å moves close to the onset of the water oxygen density peak at about 2 Å, which leads to a less ordered water multilayer, as illustrated in the broader Ow and Hw profiles. D

DOI: 10.1021/acs.jpcb.7b11697 J. Phys. Chem. B XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry B

layer slab surface (Figure 4, bottom right). As observed earlier in ref 8, using a 4 layered slab, the water molecule spontaneously dissociates and transfers one proton to the titania surface. Similarly, the mode of water adsorption on the nanotube surface was also investigated from single molecule to monolayer coverage. At submonolayer coverage, static geometry optimizations were performed. The starting configurations were generated by systematically constructing low energy structures, considering molecular adsorption of intact water molecules, as well as dissociation into an OH group bound at Ti sites and a hydrogen adsorbed at a surface oxygen site. Figure 5 shows the

Figure 3. Radial distribution functions, gTi−O(r), between surface Ti and water oxygen atoms (top) and gO−H between surface oxygen atoms and water hydrogen atoms on defect-free (red) and S-doped (blue) nanotube surfaces. Lines with symbols are for 1 ML coverage, full and dashed lines are for 2 ML water coverage.

Figure 5. Most stable adsorption configurations on the nanotube slab surface for molecular (M) and dissociative (D) adsorption at submonolayer coverages of 1/4 ML and 1/2 ML.

most stable adsorption configurations at 1/4 ML and 1/2 ML for molecular and dissociative adsorption. In the regime of isolated molecules (1/4 ML) and at 1/2 ML coverage corresponding to 1 and 2 water molecules in a (2 × 2) supercell, respectively, molecular adsorption is preferred. The metastable structures can be found in the Supporting Information. To understand the factors favoring molecular adsorption over dissociative adsorption, we further analyzed contributions to the binding energies of the most stable structures (Table 1), following the decomposition scheme described in a recent study.31 The corresponding values of these contributions for the 2D flat surface are also reported for comparison. For dissociated water molecules with O−H distances larger than 1.6 Å, spin polarization was taken into account in the evaluation of the water relaxation energies. Due to the overbinding character of PBE-D3, the bond dissociation energy of 5.99 eV slightly overestimates the experimental value and that of high-level calculations (5.46 and 5.29 eV, respectively). The large distance between the dissociated hydrogen atom and the OH-group (3.3 Å) results in very high water relaxation energies (5.99 eV).

Figure 4. Characteristic snapshots of water at 1 ML and 1/4 ML coverage, adsorbed on a 4-layered and a 3-layered anatase TiO2 (001) 2D slab surface, respectively.

conclude that neither nanostructuring (which here corresponds to curvature effects and the reduced (3 layer) slab thickness) nor solvation (water coverage) alters the interface chemistry. A similar observation also applies at low coverage. These conditions we investigated by simulating the adsorption of a single water molecule in a (2 × 2) supercell using a 3 TiO2 E

DOI: 10.1021/acs.jpcb.7b11697 J. Phys. Chem. B XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry B Table 1. Binding Energy Ebin Decomposition of the Most Stable Adsorption Configurations on the Nanotube Slab Surface for Molecular (M) and Dissociative (D) Adsorption at Submonolayer Coverage (1/4 ML and 1/2 ML) in Four Contributions: (i) Surface/Water Interaction Energy E(TiO2/H2O), (ii) Water/Water Interaction Energy E(H2O/ H2O), (iii) Surface Relaxation Energy E(TiO2), and (iv) Water Relaxation Energy E(H2O)a (2 × 2)

supercell coverage (ML)

surface in the case of dissociation requires large modifications in the substrate geometry reflected by the high relaxation energy, E(TiO2) = 2.07 eV, compared to only 0.01 eV for molecular adsorption. Thus, the gain in water−surface interaction energy due to dissociation (−1.65 eV) is less than the increase in relaxation energies required to form the dissociated structure (2.06 eV); therefore, the dissociation of an isolated water molecule is not favorable on the TiO2(001) nanotube surface. Meanwhile, on the flat surface where spontaneous dissociation of single molecules is observed (Figure 4, bottom left), less unfavorable adsorption-induced substrate relaxations (1.48 eV vs 2.07 eV) occur. The substrate relaxations due to dissociation on the nanotube correspond to a 0.18 Å extension of the Ti−Ti distances along the nanotube axis and a 0.14 Å contraction of the same distances along the corresponding direction for dissociative adsorption on the flat surface. Similar trends are observed at 1/2 ML. For thin water films with 1 ML coverage on the nanotube surface modeled by a (3 × 3) supercell (see above), the large number of possible structures and the complexity of the Hbonding networks preclude a rational construction of all possible arrangements for water. Therefore, snapshots of MD runs were taken at regular time intervals (after each 1 ps) and subjected to geometry optimization. We found that molecular adsorption still prevails, as no dissociation is observed. The absence of dissociative adsorption on the nanotube surface holds for our slab model, as well as for the original nanotubes with (0,36) chirality. Based on these observations, it can be concluded that molecular adsorption on the nanotube is most likely a curvature effect due to strain-induced changes in the electronic structure of the slab, which leads to weaker interactions, which move the equilibrium position of the molecules slightly away from the nanotube surface relative to the unstrained 2D TiO2/water interface. The binding energies of equilibrium configurations at 1 ML coverage are found to be

1/4

1/4

1/4c

1/2

1/2

n (H2O)

1

1

1

2

2

dissociation

M

D

D

M

D

E(TiO2/H2O) E(H2O/H2O) E(TiO2) E(H2O) Ebinb

−0.85 −0.01 0.01 0.13 −0.70

−8.36 −0.01 2.07 5.99 −0.29

−8.61 −0.01 1.48 5.92 −1.22

−0.69 −0.06 0.005 0.065 −0.68

−8.04 −0.08 1.78 5.99 −0.35

a

All energies are in eV. bBinding energy per molecule: at 1/2 ML the water relaxation energies are identical, and the average is reported. c Flat unstrained 2D surface.

From Table 1 it can be seen that at 1/4 ML, the energy corresponding to adsorbing and dissociating a relaxed gas-phase water molecule on the frozen TiO2 surface (neglecting water− water interactions) can be obtained by combining the interaction energy of water with the surface E(TiO2/H2O) and the relaxation energy of water E(H2O). The resulting value for the isolated dissociated water molecule (−2.37 eV) emphasizes the high strength of this adsorbate−surface interaction via one Ti5c−OH bond and one O2c−H bond. The corresponding value for undissociated water, which is adsorbed via one Ti5c−O bond and two weak H-bonds to the surface, is −0.72 eV. The very strong interaction with the

Figure 6. Selected radial distribution functions for N-doped and N+S-codoped nanotubes at 1 and 2 ML water coverage. F

DOI: 10.1021/acs.jpcb.7b11697 J. Phys. Chem. B XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry B in the range of −0.7 to −0.8 eV, showing that only weakly bound intact water molecules are present on the nanotube surface. These findings are consistent with recent NMR measurements,32 which also concluded that on water covered TiO2 (001) nanotubes, only intact water molecules are present. Doping with nitrogen promotes proton transfer to the surface. After introducing an N dopant on the surface (replacing one O2c atom) and adding an extra H atom inside the water film, proton transfer takes place from the hydronium ion to the surface N atom after 0.5 ps. This is obvious from the N-Hw RDF (Figure 6, top left), where the first maximum is centered at the characteristic NH bond distance of 1 Å. The running integration numbers n(r) of the RDF in Figure 6 (top right), with their value of ≈1 or larger at r = 1.5 Å, show that this process in all cases is complete and irreversible. At 2 ML water coverage and in the absence of sulfur, a second proton transfer from one of the water molecules takes place to form a NH+2 group (the N−H running integral is 2 at 1.5 Å in this case) and an OH− ion in the first water layer, which binds strongly to a Ti5c surface atom, as illustrated by the peak at 1.9 Å in Figure 6 (bottom left). These observations are consistent with FTIR spectral profiles of N-doped TiO2 nanocatalysts prepared by the solvothermal approach, which were reported in ref 33. There, the authors observed a peak at around 1630 cm−1, which they assigned to the bending vibrations of O−H and N−H. In addition, a broad band at 3400 cm−1, corresponding to surface hydroxyl groups and absorbed water molecule, was observed.

Figure 7 shows the larger exclusion volume around the S atom, as can be seen from a comparison with gO2c−Ow in Figure 6 (bottom right). Water in the vicinity of the S site interacts only weakly with the sulfur atoms and thus with the surface, as can be seen from the shift of the first maximum from the O−H RDF at 1.8−1.9 Å to the S−H RDF at 2.2−2.3 Å. Consistent with the larger distance, the height of the S−H maximum is smaller than that of the O−H and the N−H maxima in Figures 3 and 6 on the defect-free and N-doped surfaces, respectively. Thus, S-doping weakens hydrogen bonding from water to the surface. This shows that weakening the water−surface interaction due to the presence of a S dopant is not a local effect, and it could explain why in the case of N- and Scodoping, the surface-assisted proton transfer observed on the N-doped surface at 2 ML does not take place, and no OH groups are present. The weaker interactions between water and the S-doped TiO2 can be regarded as an indication for low reactivity compared to N-doped TiO2. Such a low reactivity of S-doped anatase (001) surfaces was also reported in an experiment carried out by Gang et al.,34 where it is seen that even under visible light irradiation, these surfaces contain only a small concentration of OH groups compared to the N-doped (001) surfaces. The results reported above also allow a qualitative understanding of the curvature-induced superhydrophilicity of anatase TiO2 nanotubes, as reported in many studies.18−20 Balur et al.18 showed that the formation of nanotubular TiO2 layers from the originally flat hydrophilic TiO2 surface significantly changes the surface wetting properties, which become superhydrophilic. Sumita et al.9 showed that the flat TiO2 (001) surface interacts with water only via the formation of surface OH− groups and a few molecularly adsorbed water molecules on Ti sites, while many of the undercoordinated surface atoms do not interact with water. However, our investigations showed that water forms strong H bonds to these undercordinated surface O sites in addition to being molecularly adsorbed on Ti sites, as shown in Figure 3 (bottom) and Figure 6 (bottom right). Even in the case of Sdoping, where the interaction with the surface is weakened, H bonds to the surface are still formed in the vicinity of the S site. This agrees with the complete (superhydrophilic) spreading of water on the nanotube surface, as observed in ref 18.



CONCLUSION

Our molecular dynamics simulations show that on the defectfree anatase TiO2 (001) nanotube surface water adsorbs molecularly via weak interactions with the Ti sites and hydrogen bonds to surface oxygens. This binding mode is a consequence of strain-induced curvature effects, as our investigations have shown that neither the nanostructuring by slab thickness reduction nor solvation of thin planar TiO2 (001) films alters the surface chemistry of the partially dissociated contact layer. The water molecules form a relatively strong hydrogen bond network, both for a buckled monolayer and for a multilayer-like (2 ML) film. While doping with sulfur, at the atom fraction (1/18) studied, weakens the interactions between the surface and water, nitrogen doping renders the surface more reactive to water, with a proton transfer to the surface and the formation of an NH group at the N site taking place in every studied case. At 2 ML coverage, even a second surface-assisted proton transfer takes place within the water film, resulting in the formation of an OH− group and an NH+2

Figure 7. Radial distribution functions (top) and running integrals (bottom) for S-doped and N+S-codoped nanotube surfaces. G

DOI: 10.1021/acs.jpcb.7b11697 J. Phys. Chem. B XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry B

(10) Ferrari, A. M.; Szieberth, D.; Zicovich-Wilson, C. M.; Demichelis, R. Anatase(001) 3 ML Nanotubes, The First TiO2 Nanotube With Negative Strain Energies: A DFT Prediction. J. Phys. Chem. Lett. 2010, 1, 2854−2857. (11) Piskunov, S.; Lisovski, O.; Begens, J.; Bocharov, D.; Zhukovskii, Y. F.; Wessel, M.; Spohr, E. C-, N-, S-, and Fe-Doped TiO2 and SrTiO3 Nanotubes for Visible-Light-Driven Photocatalytic Water Splitting: Prediction from First Principles. J. Phys. Chem. C 2015, 119, 18686−18696. (12) Roy, P.; Kim, D.; Lee, K.; Spiecker, E.; Schmuki, P. TiO2 Nanotubes and Their Application in Dye-Sensitized Solar Cells. Nanoscale 2010, 2, 45−59. (13) Lisovski, O.; Kenmoe, S.; Piskunov, S.; Bocharov, D.; Zhukovskii, Y. F.; Spohr, E. Validation of a Constrained 2D Slab Model for Water Adsorption Simulation on 1D Periodic TiO2 Nanotubes. Comput. Condens. Matter, in press 10.1016/ j.cocom.2017.11.004. (14) Casarin, M.; Vittadini, A.; Selloni, A. First Principles Study of Hydrated/Hydroxylated TiO2 Nanolayers: From Isolated Sheets to Stacks and Tubes. ACS Nano 2009, 3, 317−324. PMID: 19236066 (15) Yang, H. G.; Sun, C. H.; Qiao, S. Z.; Zou, J.; Liu, G.; Smith, S. C.; Cheng, H. M.; Lu, G. Q. Anatase TiO2 Single Crystals with a Large Percentage of Reactive Facets. Nature 2008, 453, 638−641. (16) Jung, M.-H.; Ko, K. C.; Lee, J. Y. Single Crystalline-Like TiO2 Nanotube Fabrication with Dominant (001) Facets Using Poly(vinylpyrrolidone) for High Efficiency Solar Cells. J. Phys. Chem. C 2014, 118, 17306−17317. (17) Ge, M.; Li, Q.; Cao, C.; Huang, J.; Li, S.; Zhang, S.; Chen, Z.; Zhang, K.; Al-Deyab, S. S.; Lai, Y. One-dimensional TiO2 Nanotube Photocatalysts for Solar Water Splitting. Adv. Sci. 2017, 4, 1600152. (18) Balaur, E.; Macak, J. M.; Tsuchiya, H.; Schmuki, P. Wetting Behaviour of Layers of TiO2 Nanotubes with Different Diameters. J. Mater. Chem. 2005, 15, 4488−4491. (19) Prakash, C. J.; Raj, C. C.; Prasanth, R. Fabrication of Zero Contact Angle Ultra-Super Hydrophilic Surfaces. J. Colloid Interface Sci. 2017, 496, 300−310. (20) Ghicov, A.; Schmuki, P. Self-ordering Electrochemistry: A Review on Growth and Functionality of TiO2 Nanotubes and Other Self-Aligned MOx Structures. Chem. Commun. 2009, 20, 2791−2808. (21) Shin, D. H.; Shokuhfar, T.; Choi, C. K.; Lee, S.-H.; Friedrich, C. Wettability Changes of TiO2 Nanotube Surfaces. Nanotechnology 2011, 22, 315704. (22) Kowalski, P. M.; Meyer, B.; Marx, D. Composition, Structure, and Stability of the Rutile TiO2(110) Surface: Oxygen Depletion, Hydroxylation, Hydrogen Migration, and Water Adsorption. Phys. Rev. B: Condens. Matter Mater. Phys. 2009, 79, 115410. (23) Hutter, J.; Iannuzzi, M.; Schiffmann, F.; VandeVondele, J. CP2K: Atomistic Simulations of Condensed Matter Systems. Wiley Interdiscip. Rev.: Comput. Mol.Sci. 2014, 4, 15−25. (24) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868. (25) Gillan, M. J.; Alfè, D.; Michaelides, A. Perspective: How Good is DFT for Water? J. Chem. Phys. 2016, 144, 130901. (26) Gaiduk, A. P.; Gygi, F.; Galli, G. Density and Compressibility of Liquid Water and Ice from First-Principles Simulations with Hybrid Functionals. J. Phys. Chem. Lett. 2015, 6, 2902−2908. (27) Kumar, N.; Kent, P. R. C.; Wesolowski, D. J.; Kubicki, J. D. Modeling Water Adsorption on Rutile (110) Using van der Waals Density Functional and DFT+U Methods. J. Phys. Chem. C 2013, 117, 23638−23644. (28) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. (29) Selli, D.; Fazio, G.; Seifert, G.; Di Valentin, C. Water Multilayers on TiO2(101) Anatase Surface: Assessment of a DFTB-Based Method. J. Chem. Theory Comput. 2017, 13, 3862−3873.

ion on the surface. This effect is present only in the absence of sulfur codoping, due to the generally weaker water−nanotube interactions. The consequences of this behavior for the density of excited states will have to be investigated in a forthcoming publication.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcb.7b11697. Supercell representation of the pristine nanotube and metastable adsorption configurations on the nanotube slab surface for molecular and dissociative adsorption at 1/4ML and 1/2ML (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Stephane Kenmoe: 0000-0003-3622-2716 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This study was supported by the ERA.Net RUS Plus project 237 Watersplit. The authors gratefully acknowledge computing time granted by the Center for Computational Sciences and Simulation (CCSS) of the Universität of Duisburg-Essen and provided on the supercomputer magnitUDE (DFG Grants INST 20876/209-1 FUGG and INST 20876/243-1 FUGG) at the Zentrum für Informations und Mediendienste (ZIM). E.S. is also grateful for support by the Cluster of Excellence RESOLV (EXC1069) funded by the Deutsche Forschungsgemeinschaft.



REFERENCES

(1) Kudo, A.; Miseki, Y. Heterogeneous Photocatalyst Materials for Water Splitting. Chem. Soc. Rev. 2009, 38, 253−278. (2) Kavan, L.; Grätzel, M.; Gilbert, S. E.; Klemenz, C.; Scheel, H. J. Electrochemical and Photoelectrochemical Investigation of SingleCrystal Anatase. J. Am. Chem. Soc. 1996, 118, 6716−6723. (3) Sun, C.; Liu, L.-M.; Selloni, A.; Lu, G. Q. M.; Smith, S. C. TitaniaWater Interactions: A Review of Theoretical Studies. J. Mater. Chem. 2010, 20, 10319−10334. (4) Ziółkowski, J. New Method of Calculation of the Surface Enthalpy of Solids. Surf. Sci. 1989, 209, 536−561. (5) Lazzeri, M.; Vittadini, A.; Selloni, A. Structure and Energetics of Stoichiometric TiO2 Anatase Surfaces. Phys. Rev. B: Condens. Matter Mater. Phys. 2001, 63, 155409. (6) Hosseinpour, S.; Tang, F.; Wang, F.; Livingstone, R. A.; Schlegel, S. J.; Ohto, T.; Bonn, M.; Nagata, Y.; Backus, E. H. G. Chemisorbed and Physisorbed Water at the TiO2/Water Interface. J. Phys. Chem. Lett. 2017, 8, 2195−2199. (7) Agosta, L.; Brandt, E. G.; Lyubartsev, A. P. Diffusion and Reaction Pathways of Water Near Fully Hydrated TiO2 Surfaces from Ab Initio Molecular Dynamics. J. Chem. Phys. 2017, 147, 024704. (8) Vittadini, A.; Selloni, A.; Rotzinger, F. P.; Grätzel, M. Structure and Energetics of Water Adsorbed at TiO2 Anatase 101 and 001 Surfaces. Phys. Rev. Lett. 1998, 81, 2954−2957. (9) Sumita, M.; Hu, C.; Tateyama, Y. Interface Water on TiO2 Anatase (101) and (001) Surfaces: First-Principles Study With TiO2 Slabs Dipped in Bulk Water. J. Phys. Chem. C 2010, 114, 18529− 18537. H

DOI: 10.1021/acs.jpcb.7b11697 J. Phys. Chem. B XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry B (30) Marx, D. Proton Transfer 200 Years after von Grotthuss: Insights from Ab Initio Simulations. ChemPhysChem 2006, 7, 1848− 1870. (31) Kenmoe, S.; Biedermann, P. U. Water Aggregation and Dissociation on the ZnO(10−10) Surface. Phys. Chem. Chem. Phys. 2017, 19, 1466−1486. (32) Mogilevsky, G.; Chen, Q.; Kulkarni, H.; Kleinhammes, A.; Mullins, W. M.; Wu, Y. Layered Nanostructures of Delaminated Anatase: Nanosheets and Nanotubes. J. Phys. Chem. C 2008, 112, 3239−3246. (33) Yang, G.; Jiang, Z.; Shi, H.; Xiao, T.; Yan, Z. Preparation of Highly Visible-Light Active N-doped TiO2 Photocatalyst. J. Mater. Chem. 2010, 20, 5301−5309. (34) Liu, G.; Sun, C.; Smith, S. C.; Wang, L.; Lu, G. Q. M.; Cheng, H.-M. Sulfur Doped Anatase TiO2 Single Crystals with a High Percentage of 001 Facets. J. Colloid Interface Sci. 2010, 349, 477−483.

I

DOI: 10.1021/acs.jpcb.7b11697 J. Phys. Chem. B XXXX, XXX, XXX−XXX