Next Article in Journal
Optical-Thermally Excited Graphene Resonant Mass Detection: A Molecular Dynamics Analysis
Previous Article in Journal
Bimetallic Nanocrystals: Structure, Controllable Synthesis and Applications in Catalysis, Energy and Sensing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

2D Slab Models of Nanotubes Based on Tetragonal TiO2 Structures: Validation over a Diameter Range

1
Institute of Solid State Physics, University of Latvia, LV-1063 Riga, Latvia
2
Transport and Telecommunication Institute, LV-1019 Riga, Latvia
3
Department of Theoretical Chemistry, University of Duisburg-Essen, D-45141 Essen, Germany
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(8), 1925; https://doi.org/10.3390/nano11081925
Submission received: 30 June 2021 / Revised: 20 July 2021 / Accepted: 21 July 2021 / Published: 26 July 2021

Abstract

:
One-dimensional nanomaterials receive much attention thanks to their advantageous properties compared to simple, bulk materials. A particular application of 1D nanomaterials is photocatalytic hydrogen generation from water. Such materials are studied not only experimentally, but also computationally. The bottleneck in computations is insufficient computational power to access realistic systems, especially with water or another adsorbed species, using computationally expensive methods, such as ab initio MD. Still, such calculations are necessary for an in-depth understanding of many processes, while the available approximations and simplifications are either not precise or system-dependent. Two-dimensional models as an approximation for TiO2 nanotubes with (101) and (001) structures were proposed by our group for the first time in Comput. Condens. Matter journal in 2018. They were developed at the inexpensive DFT theory level. The principle was to adopt lattice constants from an NT with a specific diameter and keep them fixed in the 2D model optimization, with geometry modifications for one of the models. Our previous work was limited to studying one configuration of a nanotube per 2D model. In this article one of the models was chosen and tested for four different configurations of TiO2 nanotubes: (101) (n,0), (101) (0,n), (001) (n,0), and (001) (0,n). All of them are 6-layered and have rectangular unit cells of tetragonal anatase form. Results of the current study show that the proposed 2D model is indeed universally applicable for different nanotube configurations so that it can be useful in facilitating computationally costly calculations of large systems with adsorbates.

1. Introduction

Performing quantum-chemical calculations of large systems, e.g., nanotubes (NTs) of realistic diameters with some chemical molecule species (e.g., water) adsorbed on one or both wall sides with computationally-costly methods such as ab initio Molecular Dynamics is one of the current problems in computational chemistry. The systems which modern supercomputers are able to simulate within a reasonable time are usually too small to match the finest systems accessible in experiments. Therefore, some computational shortcuts were developed to bridge this gap. Among others, there is an established approach to use 2D structures as an approximation to 1D NTs.
Other researchers have offered shortcuts for simulation of mechanical properties [1,2] and electronic properties [3,4] of NTs with a non-periodic cluster model constructed of an NT segment. The drawback of the method is unphysical effects on the model boundaries which results in inaccuracies of electronic structure reproduction. There is also another, zone-folding approach [5,6,7], which allows thermodynamic NT properties. This approach benefits from exploiting model periodicity but is not capable of taking NT curvature into account. Hence, there is still a necessity for a 2D-periodic model able to describe the electronic properties of NTs.
Previously, authors of the present paper have proposed two working models of water adsorption simulation on a surface of a TiO2 nanotube via constrained slabs [8]. This included a so-called Fixed Volume Slab (FVS) model which was based on slab lattice parameter modification with respect to NT constants, and a so-called Constrained Fixed Volume Slab (CFVS) model which is based not on a regular TiO2 slab with modified lattice constants, but on a 2D structure with internal coordinates based on nanotubular unit cell geometry, and with keeping coordinates of one of the TiO2 units fixed. The model construction was performed at the DFT level for further use at the ab initio MD level. More details are available in [8].
The aforementioned models have already been successfully used for their purpose, ab initio MD simulations of water adsorption on the surface of TiO2 NTs. This included a study of water adsorption on 2D-representation of 9-layered TiO2 nanotube, on clean and doped surfaces [9]. The CFVS model was also used to study the structure, energetics, and the photoelectrochemical oxidation of water on the aforementioned structures by means of the DFT + U method and a plane-wave-based computational code Quantum Espresso [10,11]. The results of these studies are summarized in a review [12].
Further on, the 2D slab models were validated over a diameter range: FVS—against (101) 6-layered (0,n) NTs (adsorption on the inner surface) with n ranging from 8 to 50 and diameter—from 2.85 to 16.50 nm, and CFVS—against (001) 9-layered (n,0) NTs with n in the range from 8 to 50 and diameter—from 1.9 to 6.32 nm. Such breadth allowed to cover all necessary domains: small NTs with prominent curvature effects, medium NTs, and large NTs with properties converging to a regular TiO2 slab. Configuration of an NT wall made each of the models more suitable for a specific NT type, please refer to our previous work for a detailed explanation [8,13].
The next step of 2D model development, considered in the present article, is dedicated to validating a single slab model against different configurations of TiO2 NTs. In the previous work, each of the 2D models was validated against a diameter range of one specific NT configuration. In turn, validation of a single 2D model against different NT configurations would create a more solid fundament for concluding on the model’s universality. Since the FVS model is less time-consuming in terms of construction, in the present study it was decided to limit the scope to the FVS model and TiO2 NTs of four different configurations: (101) 6-layered (n,0) and (0,n), and (001) 6-layered (n,0) and (0,n), with n ranging from 12 to 40 for each NT with an increment of 4. In every case, only adsorption on the outer surface is studied since it makes the most prominent contribution to the water adsorption process. The validation criteria are single water molecule adsorption energy and band gap maximum/minimum positions calculated for structures without water adsorbed.
Obviously, water adsorption causes differences in band edge positions. Still, it is not the task of the current paper to study the effects of water adsorption, since more advanced approaches and models are needed for this. Instead, here the comparison is performed between the full-size NTs and their 2D models, and the most important is to consistently compare equivalent parameters, i.e., between two systems with water, or between two systems without water.
The general motivation for using NTs and TiO2 as the material of the study is as follows.
Among many functional nanomaterials, nanostructured TiO2 are of great interest due to their unique properties and numerous practical applications [14,15,16,17]. The main interest in titania-based nanomaterials is basically associated with such high-efficient applications as lithium-ion batteries, solar cells, gas sensors, supercapacitors, etc., [18,19,20,21,22,23]. Moreover, active investigations are related to the photocatalytic activity of titania-based materials, including nanopowders and thin films [24,25,26,27,28,29]. Due to chemical stability, non-toxicity, low cost, and high availability, titanium dioxide is considered the most promising photocatalyst for the degradation of organic pollutants in water and air, as well as for water splitting and hydrogen production [30,31,32]. The additional importance of titania is related to its applications as protective coatings in microelectronic and optical devices and as luminescent compounds [33,34,35,36,37,38].
Titanium dioxide was the pioneering material in the field of photocatalytic water splitting for hydrogen generation. For this purpose, it was used for the first time in 1972 by Honda and Fujishima [39], and in the following 5 decades, a lot of attention was attracted to this material due to its prospective characteristics, such as its plenitude, low costs, non-toxicity, efficient separation and migration of charges. In addition, the position of the valence band maximum (VBM) is well-matching to hydroxyl group oxidation potential, thus providing suitable conditions for this half-reaction. On the other hand, the bandgap of TiO2 is wide (~3.2 eV in the form of anatase), and, consequently, the position of the conduction band minimum (CBM) is too low with respect to hydrogen cation reduction potential [40]. For efficient photocatalysis, gaps in the range of 1.5–2.5 eV are usually mentioned as optimal, hence a necessity to modify the material arises.
One of the ways of improving photocatalyst efficiency is the reduction of its dimensionality, to 2D and further on to 1D materials, since lower dimensionality provides shorter distances of charge migration and more efficient charge separation. The photocatalytic potential of nanotubes, nanorods, and nanowires is frequently studied. The favorable effect of low dimensions is especially prominent in NTs due to their hollow structure [41,42].
A particular problem of lowering dimensionality of TiO2 structures is the increase in the bandgap. One of the common ways to overcome this issue is to use dopant atoms in order to narrow the bandgap. This was a particular subject of interest for our group, and a series of studies have been completed. The last publication of this series [43] resulted in recommendations on doping TiO2 NTs in order to narrow the bandgap and align valence band maximum and conduction band minimum with respect to corresponding redox potentials. Carbon, nitrogen, sulphur and iron atom defects were studied, and the study concluded with a recommendation to combine S and N defects for the optimal band edge alignment.
While NTs are successfully fabricated from various materials, including titania [41], their computational studies are not abundant, mostly because the simulation of realistic NTs and other 1D nanostructures is too costly. The diameter of the finest synthesized TiO2 NTs is on the order of 10 nm, and it implies thousands of atoms in a computational model. This is still accessible by some computational codes, such as CRYSTAL [44], when it is possible to use rototranslational symmetry in order to cut the required computational resources. Thus, when there are some chemical species in the disordered phase adsorbed on such a structure, symmetry exploitation becomes inapplicable, and other solutions are needed. This closes the loop of discussion back to the motivation of developing 2D slab models of nanotubes.

2. Computational Details

Two-dimensional FVS models of the aforementioned four configurations of TiO2 NTs were developed at the DFT theory level, using hybrid exchange functional B3LYP and LCAO basis sets with localized Gaussian-type functions. The periodic code CRYSTAL [44] developed by Italian scientists at Torino university was used. In the program, there is a possibility to use rototranslational symmetry to essentially decrease needed computational resources.
In brief, the B3LYP functional was used with altered non-local HF exchange contribution (14% instead of the default 20%). The following basis sets were used: for Ti—efficient core potential basis set developed by Hay and Wadt [45] with 411sp−311d configuration for the outer shell, for O—a full-electron basis set with 6s−311sp−1d configuration [46], for H—a TZVP basis set [47]. The reasoning consisted in finding a tandem of functional and basis sets that ensure the best match of VBM and CBM to the experimental data. For the explicit procedure of basis set and exchange-correlation functional choice, the reader is kindly referred to our previous work [43].
In our previous work [8,13], the 6-layered (101) NTs and 9-layered (001) NTs had a prominent difference in the wall thickness and, consequently, radial distances until every non-equivalent atom. In the present study, both (101) and (001) NTs are 6-layered, which makes their wall thickness more similar, although the (101) configuration is thinner, see Figure 1 and Figure 2 below. In Table 1 and Table 2 main structure parameters of the (101) NTs with (n,0) and (0,n) chirality indices are presented, and in Table 3 and Table 4—parameters of the (001) NTs, also with (n,0) and (0,n) chirality indices. The chirality index is defined as the number of elementary units replicated circumferentially in an NT-ringed unit. When there is water in the model, it implies coverage of one water molecule per unit cell. The same will be true for 2D models with water.
In Table 1 it is visible that for pristine NTs a increases along with diameter growth, with approximately the same amplitude as in the first case. The a × b product also changes while b decreases. The magnitude of change is essentially higher for the b constant. In the case of NTs with water adsorbed, the change in a constant is even less prominent, while b changes within approximately the same range both for the case with water and without. Similar considerations are applicable also to (0,n) configuration. The largest difference is in larger NT diameters due to the rectangular unit cell rotated by 90°, another difference is the interchanged a and b constants.
When comparing Table 1 and Table 2 to Table 3 and Table 4, it is visible that the general trends are similar. The difference is that in the case of (001) NTs, there is no essential difference in diameter between (n,0) and (0,n) configurations with identical n. Another important difference is that a × b product for (001) NTs is several times lower than for (101) NTs, but the magnitude for its change is higher.
In our previous studies [8,13] we dealt with two types of models—the aforementioned FVS and CFVS. In this study, we use only the FVS model (Figure 3) for multiple NT configurations, therefore we will limit ourselves to a brief illustration of its principle.
The lattice constants are taken from an NT and used in a TiO2 anatase slab instead of its original constants. It is important to follow the correct order of constants. So, in the (n,0) case a and b from an NT become a and b in the 2D model, while in the (0,n) case a and b from an NT become, in turn, b and a in the 2D model. The lattice parameters of the 2D model are fixed to the same value throughout geometry optimization. In the present study, water adsorption on the outer NT surface is investigated, so the b constant corresponding to the outer NT layer is used. There are also options to use lattice parameters from a reference NT with a clean surface, or with adsorbed water, or the average of these two options.

3. Results and Discussion

3.1. The (101) NTs with (n,0) and (0,n) Configuration

The results section for both (101) and (001) configurations begins with a discussion on full-size NT and 2D model geometries. In Figure 4 and Figure 5, geometries of reference NTs (40,0) and (0,40) are presented, with their respective 2D models. It is visible that the slab model looks similar to the original NT, although there is no curvature in the planar structure. In both cases, water adsorbs with oxygen pointing at surface titanium, but in the full-size NT models hydrogens point at two surface oxygens symmetrically, while in the 2D model one of the water hydrogens points somewhat outwards. Such a trend was in general observed for the rest diameters of NTs. Two-dimensional models of smaller nanotubes, with a chirality index n = 12–16 are more distorted, while this problem is alleviated starting from a chirality index of 20 and above. This is natural since curvature effects are prominent for smaller NTs, and in the 2D models there is no curvature accounted.
Next, the first validation criterion is discussed—the water adsorption energy. One graph per NT configuration is included. Each graph contains water adsorption energies for a series of reference NTs, shown with dark blue curves, and energies for three variations of the FVS model—with constants adopted from a NT without water (red curve), with water (yellow curve), and with constants calculated as an average between these two (green curve). A straight cyan line is added to each graph to indicate the regular slab limit for water adsorption energy. In Figure 6 results for the (101) (n,0) structures are shown.
It is seen that in the left part of the graph there are essential fluctuations. The reference curve itself does not change monotonically as the chirality index grows, there is a minimum at n = 16, which was reproduced by the FVS model variation with averaged constants. Still, the performance of 2D models in the domain of low chirality index is less accurate than for larger diameters. The largest deviation is ~0.15 eV for the reference NT and FVS model based on NTs without water. For the rest variations and neighboring chirality index the deviation is mostly around 0.05–0.10 eV.
The chaotic behavior of 2D FVS model curves stops at n = 24. Since then, the growth of all curves is rather monotonical in the direction to the slab limit (−1.04 eV), and agreement with the reference curve becomes better. It is also observed that the FVS model based on NTs without water underestimates the adsorption energy in this interval, while the variation based on NTs with water—overestimates for all cases except for n = 24. The deviation is slightly less than 0.05 eV for the red curve, and several times lower for the yellow and green ones. The 2D FVS models with averaged constants provide average results between the two aforementioned cases, which is in good agreement with our previous study [12]. They also provide somewhat better agreement than the yellow curve towards larger NT diameters.
The results section continues with Figure 7 and results for the (101) (0,n) structures.
In this case, there was no such chaotic curve behavior in the interval of small diameters for all the curves as in the case of (n,0) structures. Still, the deviation from the reference curve is again the highest for the smallest diameters, with the yellow curve being the least accurate with its ~0.08 eV deviation. Again, the curve of the FVS model with averaged constants is in between the two other models. For this curve, good agreement with the reference begins at n = 24–28, with an error of around 0.025 eV and below. The red curve crosses the reference at n = 20, but this agreement is not consistent since later on the red curve diverges from the blue one. The yellow curve continues towards larger chirality indices in parallel to the reference at around 0.04 eV distance. The reference curve and the curve of the FVS model with averaged constants both converge to the slab limit (−1.04 eV).
Next, the criterion of band edge positions will be discussed based on illustrations in Figure 8 and Figure 9.
As is seen in Figure 8, in the slab limit the CBM is slightly above 0.5 eV level, and VBM—slightly below −3.5 eV, see the cyan lines with arrows. At n = 12, all curves including the reference are approximately 0.5 eV higher than the slab limit. For the CBM, the agreement between all the curves is very good and it is difficult to distinguish between them. Still, the curve for the 2D models with averaged constants is between the other two starting from n = 24 approximately. The red curve is the closest to the reference.
For the VBM, the reference curve approaches the slab limit much faster than the others. The curves for all the three FVS models gradually approach the reference, reducing the deviation from around 0.4 eV at n = 16 to almost 0 at n = 40. Again, the red curve is closest to the reference in the interval of the larger NTs, and the green curve is in the middle.
For the (101) (0,n) structures the situation is similar, with some minor differences. The curves also start above the slab level, this time with a somewhat lower difference of around 0.25–0.4 eV. All the curves converge to the slab limit in the same way, the convergence is more prominent for the VBM. Agreement between the different curves is very good.

3.2. The (001) NTs with (n,0) and (0,n) Configuration

The section of results for the (001) structures also begins with a discussion on geometry. In Figure 10 and Figure 11 fragments of (001) NTs (40,0) and (0,40) with water are shown, together with their 2D FVS models. It is possible to see the similarity between the corresponding structures, although there is no curvature in the 2D models. In the case of the (001) NTs, also water orientation agreement is better. It is visible that not only the water oxygen but also both hydrogen atoms have similar positions in both (n,0) and (0,n) cases. The general trend is that the 2D structure keeps adequate similarity to the reference NT down to chirality indices of 20–16.
Water adsorption energies are shown for (001) (n,0) structures in Figure 12. Again, there is a curve for the reference NTs, a curve for each of the three FVS model variations, and the regular slab limit is also marked, which is −0.72 eV in the (001) case.
As is seen in Figure 12, the reference curve grows rather monotonically, with a single exception of the peak at n = 28, and converges to the slab limit. In turn, the FVS model curves exhibit chaotic behavior in the interval of lower chirality indices (12–20), sometimes approaching the reference curve accidentally. Solid agreement between the reference and the FVS model curves begins at about n = 20–24. There, the maximum deviation is lower than 0.4 eV, and it becomes negligible moving forward to higher n values. It is difficult to distinguish the best-performing FVS model variation until n = 32. However, when n > 32 the FVS model variation based on averaged constants performs better than the other two.
Figure 13 shows that the reference curve growth monotonically in the interval n = 12–32, then there is a jump between 32–36 and a plateau-like interval of 36–40. There is no chaotic curve behavior in the domain of low chirality indices. Still, the FVS curves all start with an offset of around 0.075 eV with respect to the reference curve. In the interval n = 12 –32, this offset gradually diminishes until n = 32, where the curve for FVS models based on NTs with water overlaps with the reference curve, and then diverges somewhat between n = 36 and 40. For the other two slab model variations the offset grows to 0.05 eV at n = 36, and then the curves begin to approach the reference again.
Next, the criterion of band edge positions will be discussed based on Figure 14 and Figure 15.
The slab limits are slightly above 0 eV and below −4.0 eV for CBM and VBM, respectively. The curve for the reference NTs does not fluctuate essentially, 0.5 eV at most for the CBM and even less for the VBM, returning close to the starting level when reaching chirality index 40. The 2D model curves exhibit chaotic behavior in the range of smaller chirality indices, as it has already been observed for the (n,0) structures. These curves get close to each other at n = 24–28 and further on. The deviation is large, though, around 0.5 eV for the CBM along the whole interval n = 24–40, but for the VBM the deviation decreases from around 1.0 eV at n = 28 to around 0.5 eV at n = 40.
The situation is quite different for the (001) (0,n) structures. The reference curves descend monotonically from the beginning until n = 32, then there is a steeper descent until n = 36, and a plateau until n = 40. The FVS model curves do not exhibit chaotic behavior in any range. For the CBM they begin with an offset of around 0.4 eV with respect to the reference, which then decreases to almost 0 until n = 28, and then increases again. The red curve remains the closest to the reference until n = 36, but it diverges later. The two other curves reproduce the same plateau in the interval n = 36–40.
For the VBM all the curves start from the same point at n = 12. Then the reference curve and 2D model curves follow slightly different patterns until n = 32, with the deviation reaching 0.25 eV at most. Then the patterns become similar. The deviation reaches 0.5 eV again for the curve of FVS models based on NTs without water, and stays at around 0.25 eV for the other slab model curves. Though, the red curve approaches the reference in the interval of n = 36–40.

4. Summary and Conclusions

With the results discussed above, we ensured additional validation of the 2D FVS model which was developed by us earlier for facilitation of ab initio MD simulations of water-covered NTs. In this study, the novelty in the model validation was different NT configurations as material for testing of the single model. For each of the configurations, the model performance was tested both in the domain of small NT diameters where the strain effects are high and in the interval of higher chirality indices where NT properties already approach those of a slab.
The first validation criterion was water adsorption energy. In the case of (101) NTs, the FVS model demonstrated solid reproduction of the reference data, with the largest deviation of 0.15 eV. The most accurate model in this case (the variation with averaged constants) exhibited negligible deviations at the higher chirality indices. In turn, for the (001) NTs the largest deviation was even higher than 1.0 eV for the (n,0) configuration, while for the (0,n) configuration it was always lower than 0.1 eV, i.e., negligible in the whole range. Still, even for the (n,0) configuration, the difference decreases and also becomes negligible starting from chirality indices of around 24.
The second validation criterion was positions of band edges. For the (101) structures the agreement between the reference NTs and their 2D models was satisfactory, with deviations ranging between 0.5 and 0.25 eV for the (n,0) configuration. For the (0,n) configuration the results were better, with maximum deviations around 0.25 eV and with negligible deviations towards medium and large-size NTs. In the case of the (001) (n,0) structures the agreement was weaker, especially considering the chaotic behavior, so that acceptable deviation of 0.5 eV for both gap edges simultaneously was reached only at the largest NT diameter. For the (001) (0,n) configuration, in turn, the agreement was much better, not exceeding 0.25 eV in the whole interval.
To sum up the results, the 2D FVS model in general performed well for most NTs starting from medium chirality indices, around 20–28, both in terms of water adsorption energy and band edge reproduction. In the case of bandgap edge validation for (001) (n,0) structures performance of the slab models was imprecise until the largest diameter. On the contrary, the 2D model performance was good even for lower chirality indices in multiple cases. Geometry reproduction was satisfactory for (101) NTs, and good for (001) NTs. In most cases, the FVS model based on averaged lattice constants was the most accurate, although there were exceptions.
This study has contributed with additional validation to the 2D slab models proposed by us earlier [8,13], which have the potential of facilitating computationally expensive calculations, such as ab initio MD. The novelty of this work consists in validating the same 2D FVS model against 6-layered TiO2 NTs with a rectangular unit cell and with different configurations: (101) (n,0), (101) (0,n), (001) (n,0), and (001) (0,n), and positive results were obtained for every case. This broadens the conclusions of our previous work to the statement that these models are potentially applicable to NTs of other materials built from rectangular unit cells. A topic for further studies on the construction of reduced computational models could be the performance of 2D models for NTs with hexagonal configuration, although the model construction would be more cumbersome since chirality vectors are often not parallel to the circumference and main axis in a general case of a hexagonal NT.

Author Contributions

Conceptualization, O.L.; methodology, S.P. and S.K.; software, O.L.; validation, O.L., S.P., D.B. and S.K.; formal analysis, O.L., S.P., D.B. and S.K.; investigation, O.L.; resources, S.P.; data curation, O.L.; writing—original draft preparation, O.L.; writing—review and editing, O.L., S.P., D.B. and S.K.; visualization, O.L.; supervision, S.P.; project administration, O.L., S.P. and S.K.; funding acquisition, S.P. and S.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the M-ERA.NET project “Multiscale computer modelling, synthesis and rational design of photo(electro)catalysts for efficient visible-light-driven seawater splitting” (CatWatSplit). Institute of Solid State Physics, University of Latvia as the Center of Excellence has received funding from the European Union’s Horizon 2020 Framework Program H2020-WIDESPREAD-01-2016-2017-TeamingPhase2 under Grant Agreement No. 739508, project CAMART2.

Data Availability Statement

The raw/processed data required to reproduce these findings cannot be shared at this time as the data also form a part of an ongoing study.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Lu, X.; Hu, Z. Mechanical property evaluation of single-walled carbon nanotubes by finite element modeling. Compos. Part B Eng. 2012, 43, 1902–1913. [Google Scholar] [CrossRef]
  2. Mahmoudinezhad, E.; Ansari, R.; Basti, A.; Hemmatnezhad, M. An accurate spring-mass finite element model for vibration analysis of single-walled carbon nanotubes. Comput. Mater. Sci. 2014, 85, 121–126. [Google Scholar] [CrossRef]
  3. Sodré, J.M.; Longo, E.; Taft, C.A.; Martins, J.B.L.; dos Santos, J.D. Electronic structure of GaN nanotubes. Comptes Rendus Chim. 2016, 20, 190–196. [Google Scholar] [CrossRef]
  4. Chesnokov, A.; Lisovski, O.; Bocharov, D.; Piskunov, S.; Zhukovskii, F.Y.; Wessel, M.; Spohr, E. Ab initio simulations on N and S co-doped titania nanotubes for photocatalytic applications. Phys. Scr. 2015, 90, 094013. [Google Scholar] [CrossRef] [Green Version]
  5. Wirtz, L.; Rubio, A.; De La Concha, R.; Loiseau, A. Ab initio calculations of the lattice dynamics of boron nitride nanotubes. Phys. Rev. B Condens. Matter Mater. Phys. 2003, 68, 1–13. [Google Scholar] [CrossRef] [Green Version]
  6. Popov, V.N.; Lambin, P. Radius and chirality dependence of the radial breathing mode and the G -band phonon modes of single-walled carbon nanotubes. Phys. Rev. B Condens. Matter Mater. Phys. 2006, 73, 1–9. [Google Scholar] [CrossRef]
  7. Porsev, V.V.; Bandura, A.V.; Evarestov, R.A. Temperature dependence of strain energy and thermodynamic properties of V2O5-based single-walled nanotubes: Zone-folding approach. J. Comput. Chem. 2016, 37, 1442–1450. [Google Scholar] [CrossRef]
  8. 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 2018, 15, 69–78. [Google Scholar] [CrossRef]
  9. Kenmoe, S.; Lisovski, O.; Piskunov, S.; Bocharov, D.; Zhukovskii, Y.F.; Spohr, E. Water Adsorption on Clean and Defective Anatase TiO2 (001) Nanotube Surfaces: A Surface Science Approach. J. Phys. Chem. B 2018, 122, 5432–5440. [Google Scholar] [CrossRef]
  10. Kenmoe, S.; Spohr, E. Photooxidation of Water on Pristine, S- And N-Doped TiO2(001) Nanotube Surfaces: A DFT + U Study. J. Phys. Chem. C 2019, 123, 22691–22698. [Google Scholar] [CrossRef]
  11. Kenmoe, S.; Lisovski, O.; Piskunov, S.; Zhukovskii, Y.F.; Spohr, E. Electronic and optical properties of pristine, N- and S-doped water-covered TiO2 nanotube surfaces. J. Chem. Phys. 2019, 150, 041714. [Google Scholar] [CrossRef]
  12. Kenmoe, S.; Spohr, E. Computer modeling of semiconductor nanotubes for water splitting. Curr. Opin. Electrochem. 2020, 19, 88–95. [Google Scholar] [CrossRef]
  13. Lisovski, O.; Piskunov, S.; Bocharov, D.; Kenmoe, S. 2D slab models of TiO2 nanotubes for simulation of water adsorption: Validation over a diameter range. Results Phys. 2020, 19, 103527. [Google Scholar] [CrossRef]
  14. Linitis, J.; Kalis, A.; Grinberga, L.; Kleperis, J. Photo-activity research of nano-structured TiO2 layers. IOP Conf. Ser. Mater. Sci. Eng. 2011, 23, 012010. [Google Scholar] [CrossRef] [Green Version]
  15. Quy, H.V.; Truyen, D.H.; Kim, S.; Bark, C.W. Facile Synthesis of spherical TiO2 hollow nanospheres with a diameter of 150 nm for high-performance mesoporous perovskite solar cells. Materials 2021, 14, 629. [Google Scholar] [CrossRef] [PubMed]
  16. Ramirez, L.; Ramseier Gentile, S.; Zimmermann, S.; Stoll, S. Behavior of TiO2 and CeO2 nanoparticles and polystyrene nanoplastics in bottled mineral, drinking and Lake Geneva waters. Impact of water hardness and natural organic matter on nanoparticle surface properties and aggregation. Water 2019, 11, 721. [Google Scholar] [CrossRef] [Green Version]
  17. Yermakov, A.Y.; Galakhov, V.R.; Minin, A.S.; Mesilov, V.V.; Uimin, M.A.; Kuepper, K.; Bartkowski, S.; Molochnikov, L.S.; Konev, A.S.; Gaviko, V.S.; et al. Magnetic properties, electron paramagnetic resonance, and photoelectron spectroscopy studies of nanocrystalline TiO2 co-doped with Al and Fe. Phys. Status Solidi B 2021, 258, 2000399. [Google Scholar] [CrossRef]
  18. Tsebriienko, T.; Popov, A.I. Effect of poly(titanium oxide) on the viscoelastic and thermophysical properties of interpenetrating polymer networks. Crystals 2021, 11, 794. [Google Scholar] [CrossRef]
  19. Khrunyk, Y.Y.; Belikov, S.V.; Tsurkan, M.V.; Vyalykh, I.V.; Markaryan, A.Y.; Karabanalov, M.S.; Popov, A.A.; Wysokowski, M. Surface-dependent osteoblasts response to TiO2 nanotubes of different crystallinity. Nanomaterials 2020, 10, 320. [Google Scholar] [CrossRef] [Green Version]
  20. Kutuzau, M.; Shumskaya, A.; Kaniukov, E.; Alisienok, O.; Shidlouskaya, V.; Melnikova, G.; Shemukhin, A.; Nazarov, A.; Kozlovskiy, A.; Zdorovets, M. Photocatalytically active filtration systems based on modified with titanium dioxide PET-membranes. Nucl. Inst. Methods Phys. Res. B 2019, 460, 212–215. [Google Scholar] [CrossRef]
  21. Raza, W.; Hwang, I.; Denisov, N.; Schmuki, P. Thermal ramping rate during annealing of TiO2 nanotubes greatly affects performance of photoanodes. Phys. Status Solidi A 2021, 218, 2100040. [Google Scholar] [CrossRef]
  22. Bhoomanee, C.; Sanglao, J.; Kumnorkaew, P.; Wang, T.; Lohawet, K.; Ruankham, P.; Gardchareon, A.; Wongratanaphisan, D. Hydrothermally treated TiO2 nanorods as electron transport layer in planar perovskite solar cells. Phys. Status Solidi A 2021, 218, 2000238. [Google Scholar] [CrossRef]
  23. Kia, A.M.; Bönhardt, S.; Zybell, S.; Kühnel, K.; Haufe, N.; Weinreich, W. Development of rutile titanium oxide thin films as battery material component using atomic layer deposition. Phys. Status Solidi A 2020, 217, 1800769. [Google Scholar] [CrossRef]
  24. Serga, V.; Burve, R.; Krumina, A.; Romanova, M.; Kotomin, E.A.; Popov, A.I. Extraction–pyrolytic method for TiO2 polymorphs production. Crystals 2021, 11, 431. [Google Scholar] [CrossRef]
  25. Wu, F.; Hu, X.; Fan, J.; Sun, T.; Kang, L.; Hou, W.; Zhu, C.; Liu, H. Photocatalytic activity of Ag/TiO2 nanotube arrays enhanced by surface plasmon resonance and application in hydrogen evolution by water splitting. Plasmonics 2013, 8, 501–508. [Google Scholar] [CrossRef]
  26. Regmi, C.; Lotfi, S.; Espíndola, J.C.; Fischer, K.; Schulze, A.; Schäfer, A.I. Comparison of photocatalytic membrane reactor types for the degradation of an organic molecule by TiO2-coated PES membrane. Catalysts 2020, 10, 725. [Google Scholar] [CrossRef]
  27. Serga, V.; Burve, R.; Krumina, A.; Pankratova, V.; Popov, A.I.; Pankratov, V. Study of phase composition, photocatalytic activity, and photoluminescence of TiO2 with Eu additive produced by the extraction-pyrolytic method. J. Mater. Res. Technol. 2021, 13, 2350–2360. [Google Scholar] [CrossRef]
  28. Manna, A.K.; Kanjilal, A.; Kanjilal, D.; Varma, S. Dynamic Surface evolution and scaling studies on TiO2 thin films by Ti ion implantation. Nucl. Inst. Methods Phys. Res. B 2020, 474, 68–73. [Google Scholar] [CrossRef]
  29. Nair, S.B.; John, K.A.; Joseph, J.A.; Babu, S.; Shinoj, V.K.; Remillard, S.K.; Shaji, S.; Philip, R.R. Fabrication and characterization of type-II heterostructure n:In2O3/p:in-TiO2 for enhanced photocatalytic activity. Phys. Status Solidi B 2021, 258, 2000441. [Google Scholar] [CrossRef]
  30. Li, Y.; Wang, W.; Wang, F.; Di, L.; Yang, S.; Zhu, S.; Yao, Y.; Ma, C.; Dai, B.; Yu, F. Enhanced photocatalytic degradation of organic dyes via defect-rich TiO2 prepared by dielectric barrier discharge plasma. Nanomaterials 2019, 9, 720. [Google Scholar] [CrossRef] [Green Version]
  31. Pant, B.; Park, M.; Park, S.-J. Recent advances in TiO2 films prepared by Sol-Gel methods for photocatalytic degradation of organic pollutants and antibacterial activities. Coatings 2019, 9, 613. [Google Scholar] [CrossRef] [Green Version]
  32. Zhao, W.; Yang, X.; Liu, C.; Qian, X.; Wen, Y.; Yang, Q.; Sun, T.; Chang, W.; Liu, X.; Chen, Z. Facile construction of all-solid-state Z-scheme g-C3N4/TiO2 thin film for the efficient visible-light degradation of organic pollutant. Nanomaterials 2020, 10, 600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Knoks, A.; Kleperis, J.; Grinberga, L. Raman spectral identification of phase distribution in anodic titanium dioxide coating. Proc. Est. Acad. Sci. 2017, 66, 422–429. [Google Scholar] [CrossRef]
  34. Brik, M.G.; Antic, Ž.M.; Vukovic, K.; Dramicanin, M.D. Judd-Ofelt analysis of Eu3+ emission in TiO2 anatase nanoparticles. Mater. Trans. 2015, 56, 1416–1418. [Google Scholar] [CrossRef] [Green Version]
  35. Kozlovskiy, A.; Shlimas, D.; Kenzhina, I.; Boretskiy, O.; Zdorovets, M. Study of the effect of low-energy irradiation with O2+ ions on radiation hardening and modification of the properties of thin TiO2 films. J. Inorg. Organomet. Polym. Mater. 2021, 31, 790–801. [Google Scholar] [CrossRef]
  36. Dukenbayev, K.; Kozlovskiy, A.; Kenzhina, I.; Berguzinov, A.; Zdorovets, M. Study of the effect of irradiation with Fe7+ ions on the structural properties of thin TiO2 foils. Mater. Res. Express 2019, 6, 046309. [Google Scholar] [CrossRef]
  37. Zhai, P.; Nan, S.; Xu, L.; Li, W.; Li, Z.; Hu, P.; Zeng, J.; Zhang, S.; Sun, Y.; Liu, J. Fine structure of swift heavy ion track in rutile TiO2. Nucl. Inst. Methods Phys. Res. B 2019, 457, 72–79. [Google Scholar] [CrossRef] [Green Version]
  38. Hazem, R.; Izerrouken, M.; Cheraitia, A.; Djehlane, A. Raman study of ion beam irradiation damage on nanostructured TiO2 thin film. Nucl. Inst. Methods Phys. Res. B 2019, 444, 62–67. [Google Scholar] [CrossRef]
  39. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 38, 37–38. [Google Scholar] [CrossRef]
  40. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef]
  41. Ge, M.; Cai, J.; Iocozzia, J.; Cao, C.; Huang, J.; Zhang, X.; Shen, J.; Wang, S.; Zhang, S.; Zhang, K.Q.; et al. A review of TiO2 nanostructured catalysts for sustainable H2 generation. Int. J. Hydrog. Energy 2016, 42, 8418–8449. [Google Scholar] [CrossRef]
  42. Zhukovskii, Y.F.; Pugno, N.; Popov, A.I.; Balasubramanian, C.; Bellucci, S. Influence of F centres on structural and electronicproperties of AlN single-walled nanotubes. J. Phys. Condens. Matter 2007, 19, 395021. [Google Scholar] [CrossRef] [Green Version]
  43. 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. [Google Scholar] [CrossRef]
  44. Dovesi, R.; Saunders, V.R.; Roetti, C.; Orlando, R.; Zicovich-Wilson, C.M.; Pascale, F.; Civalleri, B.; Doll, K.; Harrison, N.M.; Bush, I.J.; et al. Crystal’17 User’s Manual. 2018. Available online: https://www.crystal.unito.it/manuals/crystal17.pdf (accessed on 26 July 2021).
  45. Piskunov, S.; Heifets, E.; Eglitis, R.; Borstel, G. Bulk properties and electronic structure of SrTiO3, BaTiO3, PbTiO3 perovskites: An ab initio HF/DFT study. Comput. Mater. Sci. 2004, 29, 165–178. [Google Scholar] [CrossRef]
  46. Heyd, J.; Peralta, J.E.; Scuseria, G.E.; Martin, R.L. Energy band gaps and lattice parameters evaluated with the Heyd-Scuseria-Ernzerhof screened hybrid functional. J. Chem. Phys. 2005, 123. [Google Scholar] [CrossRef]
  47. Peintinger, M.F.; Oliveira, D.V.; Bredow, T. Consistent Gaussian basis sets of triple-zeta valence with polarization quality for solid-state calculations. J. Comput. Chem. 2013, 34, 451–459. [Google Scholar] [CrossRef]
Figure 1. Structures of 6-layered (101) TiO2 NTs. Unit cells are depicted with blue lines in the magnified segments. (a) (n,0) configuration; (b) (0,n) configuration.
Figure 1. Structures of 6-layered (101) TiO2 NTs. Unit cells are depicted with blue lines in the magnified segments. (a) (n,0) configuration; (b) (0,n) configuration.
Nanomaterials 11 01925 g001
Figure 2. Structures of 6-layered (001) TiO2 NTs. Unit cells are depicted with blue lines in the magnified segments. (a) (n,0) configuration; (b) (0,n) configuration.
Figure 2. Structures of 6-layered (001) TiO2 NTs. Unit cells are depicted with blue lines in the magnified segments. (a) (n,0) configuration; (b) (0,n) configuration.
Nanomaterials 11 01925 g002
Figure 3. Basic illustration of the 2D FVS model principle.
Figure 3. Basic illustration of the 2D FVS model principle.
Nanomaterials 11 01925 g003
Figure 4. Fragment of a 6-layered (101) TiO2 NT (40,0) and its 2D FVS model, with a multiplied unit cell.
Figure 4. Fragment of a 6-layered (101) TiO2 NT (40,0) and its 2D FVS model, with a multiplied unit cell.
Nanomaterials 11 01925 g004
Figure 5. Fragment of a 6-layered (101) TiO2 NT (0,40) and its 2D FVS model, with a multiplied unit cell.
Figure 5. Fragment of a 6-layered (101) TiO2 NT (0,40) and its 2D FVS model, with a multiplied unit cell.
Nanomaterials 11 01925 g005
Figure 6. Adsorption energies for (101) NTs (n,0) and their 2D FVS models. The cyan line with an arrow denotes the slab limit for water adsorption energy.
Figure 6. Adsorption energies for (101) NTs (n,0) and their 2D FVS models. The cyan line with an arrow denotes the slab limit for water adsorption energy.
Nanomaterials 11 01925 g006
Figure 7. Adsorption energies for (101) NTs (0,n) and their 2D FVS models. The cyan line with an arrow denotes the slab limit for water adsorption energy.
Figure 7. Adsorption energies for (101) NTs (0,n) and their 2D FVS models. The cyan line with an arrow denotes the slab limit for water adsorption energy.
Nanomaterials 11 01925 g007
Figure 8. VB top and CB bottom edge positions for (101) NTs (n,0) and their 2D FVS models. The cyan lines with arrows denote slab limit for VB maximum and CB minimum.
Figure 8. VB top and CB bottom edge positions for (101) NTs (n,0) and their 2D FVS models. The cyan lines with arrows denote slab limit for VB maximum and CB minimum.
Nanomaterials 11 01925 g008
Figure 9. VB top and CB bottom edge positions for (101) NTs (0,n) and their 2D FVS models. The cyan lines with arrows denote slab limit for VB maximum and CB minimum.
Figure 9. VB top and CB bottom edge positions for (101) NTs (0,n) and their 2D FVS models. The cyan lines with arrows denote slab limit for VB maximum and CB minimum.
Nanomaterials 11 01925 g009
Figure 10. Fragment of a 6-layered (001) TiO2 NT (40,0) and its 2D FVS model, with a multiplied unit cell.
Figure 10. Fragment of a 6-layered (001) TiO2 NT (40,0) and its 2D FVS model, with a multiplied unit cell.
Nanomaterials 11 01925 g010
Figure 11. Fragment of a 6-layered (001) TiO2 NT (0,40) and its 2D FVS model, with a multiplied unit cell.
Figure 11. Fragment of a 6-layered (001) TiO2 NT (0,40) and its 2D FVS model, with a multiplied unit cell.
Nanomaterials 11 01925 g011
Figure 12. Adsorption energies for (001) NTs (n,0) and their 2D FVS models. The cyan line with the arrow denotes the slab limit for water adsorption energy.
Figure 12. Adsorption energies for (001) NTs (n,0) and their 2D FVS models. The cyan line with the arrow denotes the slab limit for water adsorption energy.
Nanomaterials 11 01925 g012
Figure 13. Adsorption energies for (001) NTs (0,n) and their 2D FVS models. The cyan line with the arrow denotes the slab limit for water adsorption energy.
Figure 13. Adsorption energies for (001) NTs (0,n) and their 2D FVS models. The cyan line with the arrow denotes the slab limit for water adsorption energy.
Nanomaterials 11 01925 g013
Figure 14. VB top and CB bottom edge positions for (001) NTs (n,0) and their 2D FVS models. The cyan lines with arrows denote slab limits for VB maximum and CB minimum.
Figure 14. VB top and CB bottom edge positions for (001) NTs (n,0) and their 2D FVS models. The cyan lines with arrows denote slab limits for VB maximum and CB minimum.
Nanomaterials 11 01925 g014
Figure 15. VB top and CB bottom edge positions for (001) NTs (0,n) and their 2D FVS models. The cyan lines with arrows denote slab limits for VB maximum and CB minimum.
Figure 15. VB top and CB bottom edge positions for (001) NTs (0,n) and their 2D FVS models. The cyan lines with arrows denote slab limits for VB maximum and CB minimum.
Nanomaterials 11 01925 g015
Table 1. Parameters of 6-layered (101) (n,0) TiO2 NTs, simulated with and without water. The lattice parameter corresponding to NT axis is denoted with a, d stands for outer diameter, b—for outer b constant along NT circumference, and a × b—their product.
Table 1. Parameters of 6-layered (101) (n,0) TiO2 NTs, simulated with and without water. The lattice parameter corresponding to NT axis is denoted with a, d stands for outer diameter, b—for outer b constant along NT circumference, and a × b—their product.
(101) (n,0)NTs with Clean SurfaceNTs with Water Adsorbed
Chirality Index nd Outeraba × bd Outeraba × b
1216.63610.4114.35345.31916.95410.2454.43645.451
1621.04310.4144.13043.00621.54610.2494.22843.335
2025.45610.4223.99741.65126.05210.2584.09041.956
2429.88410.4253.91040.76230.56410.2643.99941.045
2834.32610.4303.84940.14935.11610.2653.93840.422
3238.75710.4353.80339.68539.66610.2673.89239.962
3643.23710.4353.77139.35244.24610.2673.85939.624
4047.67410.4393.74239.06848.82410.2693.83339.358
Table 2. Parameters of 6-layered (101) (0,n) TiO2 NTs, simulated with and without water. The lattice parameter corresponding to the NT axis is denoted with a, d stands for outer diameter, b—for outer b constant along NT circumference, and a × b—their product.
Table 2. Parameters of 6-layered (101) (0,n) TiO2 NTs, simulated with and without water. The lattice parameter corresponding to the NT axis is denoted with a, d stands for outer diameter, b—for outer b constant along NT circumference, and a × b—their product.
(101) (0,n)NTs with Clean SurfaceNTs with Water Adsorbed
Chirality Index nd Outeraba × bd Outeraba × b
1242.2883.54011.06539.16842.0863.62211.01339.885
1655.5853.53510.90838.56654.8303.61510.76038.902
2068.8773.53310.81438.21068.4213.61210.74238.800
2482.1483.53510.74837.99481.5893.61010.67538.533
2895.4983.52910.70937.79894.7553.60910.62638.346
32108.7253.53410.66937.704107.9243.60810.59038.206
36122.0953.52910.64937.580121.0943.60710.56238.098
40135.4023.52910.62937.506134.2603.60610.53938.010
Table 3. Parameters of 6-layered (001) (n,0) TiO2 NTs, simulated with and without water. The lattice parameter corresponding to the NT axis is denoted with a, d stands for outer diameter, b—for outer b constant along NT circumference, and a × b—their product.
Table 3. Parameters of 6-layered (001) (n,0) TiO2 NTs, simulated with and without water. The lattice parameter corresponding to the NT axis is denoted with a, d stands for outer diameter, b—for outer b constant along NT circumference, and a × b—their product.
(001) (n,0)NTs with Clean SurfaceNTs with Water Adsorbed
Chirality Index nd Outeraba × bd Outeraba × b
1219.8093.6015.18318.66420.0443.5835.24518.792
1622.9003.6644.49416.46824.1933.5734.74816.963
2025.2653.7433.96714.84627.2643.7844.28016.195
2428.4483.7723.72214.03930.9393.8224.04815.471
2831.9383.7823.58213.54532.4623.6953.64013.453
3235.5583.7853.48913.20636.5023.6823.58213.189
3639.1643.7873.41612.93540.2443.6913.51012.956
4042.8413.7873.36312.73643.3493.7323.40312.700
Table 4. Parameters of 6-layered (001) (0,n) TiO2 NTs, simulated with and without water. The lattice parameter corresponding to the NT axis is denoted with a, d stands for outer diameter, b—for outer b constant along NT circumference, and a × b—their product.
Table 4. Parameters of 6-layered (001) (0,n) TiO2 NTs, simulated with and without water. The lattice parameter corresponding to the NT axis is denoted with a, d stands for outer diameter, b—for outer b constant along NT circumference, and a × b—their product.
(001) (0,n)NTs with Clean SurfaceNTs with Water Adsorbed
Chirality Index nd Outeraba × bd Outeraba × b
1220.6283.2475.39817.52820.3803.2505.33317.333
1625.4683.2494.99816.24025.3233.2594.97016.194
2030.2523.2504.75015.43630.2293.2664.74615.499
2434.9723.2494.57614.86835.1043.2724.59315.028
2839.6363.2474.44514.43439.9433.2804.47914.691
3244.2953.2464.34614.10844.7743.2804.39314.410
3647.6843.0144.15912.53347.6403.0164.15512.530
4052.4613.0144.11812.41252.3683.0174.11112.404
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lisovski, O.; Piskunov, S.; Bocharov, D.; Kenmoe, S. 2D Slab Models of Nanotubes Based on Tetragonal TiO2 Structures: Validation over a Diameter Range. Nanomaterials 2021, 11, 1925. https://doi.org/10.3390/nano11081925

AMA Style

Lisovski O, Piskunov S, Bocharov D, Kenmoe S. 2D Slab Models of Nanotubes Based on Tetragonal TiO2 Structures: Validation over a Diameter Range. Nanomaterials. 2021; 11(8):1925. https://doi.org/10.3390/nano11081925

Chicago/Turabian Style

Lisovski, Oleg, Sergei Piskunov, Dmitry Bocharov, and Stephane Kenmoe. 2021. "2D Slab Models of Nanotubes Based on Tetragonal TiO2 Structures: Validation over a Diameter Range" Nanomaterials 11, no. 8: 1925. https://doi.org/10.3390/nano11081925

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop