Next Article in Journal
Effect of Electron Irradiation Fluence on InP-Based High Electron Mobility Transistors
Previous Article in Journal
A Review of THz Modulators with Dynamic Tunable Metasurfaces
Previous Article in Special Issue
Rheological Tunability of Perovskite Precursor Solutions: From Spin Coating to Inkjet Printing Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Transition Metal Substituents on Interfacial and Electronic Structure of CH3NH3PbI3/TiO2 Interface: A First-Principles Comparative Study

1
Department of Chemical and Environmental Engineering, Anyang Institute of Technology, Anyang 455000, China
2
Department of Mathematics and Physics, Anyang Institute of Technology, Anyang 455000, China
3
Integrated Composites Laboratory (ICL), Department of Chemical & Biomolecular Engineering, University of Tennessee, Knoxville, TN 37996, USA
4
Key Laboratory of Materials Processing and Mold (Zhengzhou University), Ministry of Education; National Engineering Research Center for Advanced Polymer Processing Technology, Zhengzhou University, Zhengzhou 450002, China
5
School of Urban Planning and Architecture, Southwest Minzu University, Chengdu 610041, China
6
School of Material Science and Engineering, Jiangsu University of Science and Technology, Zhenjiang 212003, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2019, 9(7), 966; https://doi.org/10.3390/nano9070966
Submission received: 21 June 2019 / Accepted: 28 June 2019 / Published: 1 July 2019
(This article belongs to the Special Issue Perovskite Materials and Devices–– Progress and Challenges)

Abstract

:
To evaluate the influence of transition metal substituents on the characteristics of CH3NH3PbI3/TiO2, we investigated the geometrical and electronic properties of transition metal-substituted CH3NH3PbI3/TiO2 by first-principles calculations. The results suggested that the substitution of Ti4+ at the five-fold coordinated (Ti5c) sites by transition metals is energetically favored. The substituted interface has enhanced visible light sensitivity and photoelectrocatalytic activity by reducing the transition energies. The transition metal substitution can effectively tune the band gap of the interface, which significantly improves the photo-reactivity. The substituted systems are expected to be more efficient in separating the photo-generated electrons-holes and active in the visible spectrum.

1. Introduction

Hybrid halide perovskites as light harvesters have been the focus of the photovoltaic field over the past years owing to their impressive power conversion efficiency (PCE) and promising commercial applications [1,2]. The CH3NH3PbI3 perovskites dominate this field and have been studied extensively [3]. The typical device architecture of perovskite solar cells (PSC) is composed of the TiO2-based electron transport layer (ETL), the perovskite (CH3NH3PbI3) absorber-based layer, the spiro-OMeTAD hole transport layer (HTL) and the corresponding electrodes [4]. The perovskite/ETL interface plays an important role in determining the charge separation and transport properties as well as the PSC device performance, which has been widely explored for many years [5,6,7,8].
TiO2 is a good candidate material due to its chemical stability, high charge transport property and low cost [9]. The electron can be effectively transported from CH3NH3PbI3 to the TiO2 layer because the conduction band of TiO2 is lower than that of CH3NH3PbI3 [10]. Generally, the ultra-thin compact TiO2 layer is prepared at a high temperature (over 450 °C) [11]. It seriously undermines the electrical properties of the ETL in conductivity, mobility, and electronic trap states, thereby affecting the efficiency and stability of PSC. Elemental substitution in the compact TiO2 layer is an effective solution to improve the electrical properties and device performance [12,13,14,15,16,17,18,19,20,21]. Transition metal substitution [22] is an effective approach to enhance the photocatalytic activity of TiO2 due to their unique d electronic configuration and spectral characteristics. According to previous studies [13,14,15,16,17,18,19,20,21], a wide range of substitutional elements such as Zn2+, Y3+, Nb5+, Ru4+ and W6+ have been investigated in TiO2. Research has made remarkable progress in identifying that the substitution of the TiO2 layer by the transition metal in PSC is an effective mean to improve the photocurrent and electron-hole recombination [14,15,16,17,18]. Substituents in TiO2 film can improve the electrical characteristics of ETL, which promote PCE and stability of PSC [13,23]. Therefore, transition metal substituents in the TiO2 layer are quite important and further characterizations are required to understand the effects of substituents in the CH3NH3PbI3/TiO2 interface. Although the electronic properties of perovskite/ETL interface have been widely investigated by experiments and density-functional-theory (DFT) calculations [24,25,26,27,28,29,30,31], the existence of theoretical studies aiming to understand the fundamental role of the interfacial substituents is still rather scarce. In addition to the primary experiments, the first-principles DFT calculations are highly important to acquire further knowledge concerning the effects of transition metal substitution and contribute to new strategies for interface optimization. The main contributions of our study are helpful to draw guidelines for substitution mechanism of the CH3NH3PbI3/TiO2 interface, thus enhancing the photovoltaic performance in PSC.

2. Methods

The Vienna ab initio simulation package (VASP) [32] was employed as the first-principles calculations platform. The computer software program is based on the DFT approach using plane wave basis within a periodic boundary condition. The projected augmented wave (PAW) [33] pseudopotentials were applied for efficient computation. The exchange and correlations items were treated within the framework of generalized-gradient approximation (GGA) of Perdew-Burke-Ernzerhof (PBE) [34]. A plane wave basis cutoff energy of 500 eV was used. Integrations in reciprocal space were sampled using the Monkhorst-Pack grids [35] with a minimum spacing of 0.2 Å−1. Convergence criteria were set as 1.0−6 eV in total energy and 1.0−2 eV/Å in atomic force, respectively. Recent theoretical studies indicated that the GW (Green’s function G with screened interaction W) and hybrid functional approach can provide an accurate description of the electronic structures [36,37]. Fortunately, DFT was able to qualitatively reproduce the GW trend. Hence, we performed GGA+U calculations on the CH3NH3PbI3/TiO2 interfaces with reasonable computational cost. Based on previous research and experiences [38,39,40,41,42,43], the GGA+U approach with the on-site Coulomb interaction correction predicted band gap correctly. The values of parameter U were 6 eV for the Ti4+ 3d orbit and 4 eV for the d-orbits of transition metal substituents. Gaussian broadening [44] with half-width of 0.1 eV for the electronic eigenvalues was used to accelerate the convergence in the k-point sum. The dipole correction was included because the interface configuration does not have mirror symmetry along the c-axis. The spin orbit coupling (SOC) effect [45] was not included because it was negligible in the geometry. The atomic structures shown were produced by using the visualization for electronic and structural analysis (VESTA) program [46].
According to the experimental results [47], there is an ordered lattice structure existing on the CH3NH3PbI3/TiO2 interface. The (110) slab of the CH3NH3PbI3 nanocrystal coordinated with the (101) slab of anatase TiO2, forming an ordered lattice structure at the interface. The lattice mismatch between CH3NH3PbI3 (110) and TiO2 (101) has been evaluated in previous studies [48]. It was found that using the experimental results of the CH3NH3PbI3 (110) surface, the band-gap only varies slightly, with a corresponding total energy decrease, suggesting that a minimal strain is introduced by the lattice mismatch between the two materials. In spite of a relatively large lattice mismatch, we selected the experimental interface [47] between CH3NH3PbI3 (110) and TiO2 (101) to carry out the investigation. To construct the aspired interface structures, the CH3NH3PbI3 (110)/TiO2 (101) interface was chosen as our objective due to the experimental results [47] and stability of the corresponding surface. The (2 × 2) supercell of the CH3NH3PbI3 (110) and the (1 × 3) supercell of the anatase TiO2 (101) were employed to decrease the misfit. The interface model employed the average size of both CH3NH3PbI3 and TiO2 slabs to make a small mismatch. The CH3NH3PbI3 perovskite was composed of the CH3NH3I and PbI2 units along the [001] direction; therefore, both CH3NH3I and PbI2 terminations using five-layer slabs were considered in this work. The anatase (101) supercell contains 36 TiO2 units or 108 substrate atoms. The interfaces were built by connecting the CH3NH3PbI3 (110) slab with the anatase (101) slab and leaving a 20 Å vacuum gap in the perpendicular direction. A schematic representation of the interfacial system used in this work is presented in Figure 1. In fact, the TiO2 was the substrate to grow perovskite films; hence, apart from the bottom two Ti and four O layers, which were maintained in their ideal bulk positions, all atomic coordinates of the others layers were fully relaxed. The transition metal-substituted anatase (101) surface model was constructed by substituting the surface Ti4+ with substituent atoms. As shown in Figure 2, there are two possible surface sites (five/six-fold coordinated Ti4+ cation, hereafter abbreviated as Ti5C and Ti6C) for the substituent atoms to replace [49]. Till date, the atomic arrangement of the CH3NH3PbI3/TiO2 is still unclear due to limit of the experimental techniques. Based on previous studies of the transition metal-substituted TiO2 surface [50,51], the interfacial configurations were carefully designed to make the substituent effect more prominent at the interface region. We assume that all six different transition metal ions partially substituted at the Ti4+ sites (Ti5C or Ti6C) correspond to the substitution concentration of 17% and the supercell is represented by Ti0.83M0.17O2 (M = Zn2+, Y3+, Zr4+, Nb5+, Ru4+, W6+). The substituents can be classified as 3d transition metal (Zn2+), 4d transition metal (Y3+, Zr4+, Nb4+, Ru4+) and 5d transition metal (W6+) ions.

3. Results and Discussions

As shown in Table 1, the optimized 0 K DFT lattice parameters of anatase TiO2 crystal are a = 3.79 Å and c = 9.53 Å, which is in agreement with previous experiments [52]. The atomic positions of tetragonal CH3NH3PbI3 are based on the results of the previous report [53,54]. The 0 K DFT lattice parameters are a = 8.80 Å and c = 13.05 Å. The strong interfacial interaction in CH3NH3PbI3/TiO2 is mainly through iodine and under-coordinated titanium atoms. The stability of the selected interface can be evaluated by comparing the calculated binding energies [55,56]. The calculated binding energies of different interfaces are listed in Table 2, together with lattice mismatch. The lattice mismatch between CH3NH3PbI3 and TiO2 was −12.0% and −13.8%, respectively. The interfacial binding energies and lattice mismatch can be predicted by the following equations [25]:
Ebinding = Eanatase + EperovskiteEtotal
Mperovskite/anatase = (aperovskiteaanatase)/aperovskite
where Etotal, Eanatase, and Eperovskite are the corresponding energies of the interface, anatase and perovskite surfaces, respectively. The aanatase and aperovskite represent the lattice parameter of TiO2 and CH3NH3PbI3, respectively. Zero energy corresponds to the energetically less-stable structure. It is not surprising that the perovskite/TiO2 interfaces without rotation are more stable than their corresponding rotated ones. This could be because of the difference in lattice mismatches. The strain may affect the interfacial stability between perovskite and TiO2. For rotation-free interfaces, the Ebinding of the two systems is quite similar. The CH3NH3+ cation interacted with TiO2 partially containing weak van der Waals (vdW) interactions. In contrast, the interaction between Pb2+ cation and TiO2 leads to the formation of stable chemical bonds. This character is similar to previous works [25,54]. The interfacial structures of the relatively stable rotation-free perovskite/TiO2 are adopted in subsequent calculations. The interface supercell lattice parameters are given by a = 10.95 Å, b = 11.64 Å and c = 49.95 Å.
To explore the influence of transition metal substitution on the interfacial stability of the CH3NH3PbI3/TiO2 interface structure, the Nb-substituted interfaces were systematically investigated. The dependence of Nb5+ substituent on the depth within the interface layers were evaluated based on the total energies (seen Table 3). The six-coordinated Ti6C substituted interface with the lowest total energy is accepted as the most stable configuration. Despite the cleaved and unsaturated bond, the five-coordinated Ti5C substituted interfaces still show relative low total energy. The Nb5+ substituent was energetically favorable at the Ti5C and Ti6C sites of the TiO2 surface. Therefore, to characterize the effect of substitution at the CH3NH3PbI3/TiO2 interface, both Ti5C and Ti6C substitution sites are considered in subsequent calculations. The substitution of Nb5+ at both Ti5C site and Ti6C sites in the top TiO2 layer was considered. For each situation, two different configurations, namely CH3NH3I/TiO2 and PbI2/TiO2, were considered in this study. The calculated interfacial binding energies and Bader charge were listed in Table 4. The interfacial binding energies become stronger after substitution, which indicates that substituting Ti4+ with Nb5+ could enhance the stability and strength of the perovskite/TiO2 interface. The CH3NH3I/TiO2 interface has a larger energy than the PbI2/TiO2 interface. The degree of charge transferring is evaluated by the Bader charge analysis [57]. The negative value means the transfer of excess electrons from the perovskite to TiO2 because the perovskite layer has a higher average potential than the TiO2 layer. As can be seen in Table 4, there is less charge transfer in the PbI2/TiO2 than that of the CH3NH3I/TiO2. This can be explained by the fact that the PbI2 layers have a relatively lower potential than the CH3NH3I layers. Moreover, it also can be seen that the Nb5+ substituents located at the Ti5C site have a larger charge transfer than that of the Ti6C site. Liu et al. reported that the potential drop on the CH3NH3I/TiO2 is deeper than that of the PbI2/TiO2 [25]. As a result, a strong accumulation region can be formed at the CH3NH3PbI3/TiO2 interface, leading to a better electron-hole separation in the PSC. To emphasize and compare the influence of more different transition metal substituents on the CH3NH3PbI3/TiO2 interface, we make the approximation that only the Ti5C-substitute interface will be considered in subsequent investigations.
Both the interfacial charge transfers and Ebinding of the transition metal-substituted CH3NH3PbI3/TiO2 interfaces in Figure 3 were combined to evaluate the influence of transition metal substitution in the PSC. Transition metals can be divided into three types: n (Nb5+, W6+), p (Zn2+, Y3+) and isovalent (Zr4+, Ru4+) substitutions. Zero energy (pristine) corresponds to the energetically less-stable structure. It can be seen that the transition metals have different Ebinding while the values of each substituted interface are positive. It also can be discerned clearly from Figure 3 that the transition metal-substituted interface has a much higher binding energy. This suggests that substituting transition metals M (M = Zn2+, Y3+, Zr4+, Nb5+, Ru4+, W6+) for Ti4+ at the interface layer could significantly enhance the interface strength between perovskite and TiO2. In addition, Figure 3 displays the comparison of charge transfers at the perovskite/TiO2 interface substituted with various transition metals. One can clearly see that the charge transfer in the interface becomes larger with the addition of transition metals. It should be pointed out that Zn2+ and Y3+ substitution for Ti4+ at the interface layer is energetically favorable in terms of binding energy (16.6 and 13.5 eV) and charge transfer (−5.8 and −3.1 e) for the PbI2/TiO2 interface. This enhancement can be attributed to the optimized energy band alignment, which could improve the electron transfer behavior between ETL and perovskite. The theoretical results can provide support for future experimental design and synthesis of a stable perovskite/TiO2 interface, possessing strong electron transfer capacity. Due to their relatively stronger binding energies, the interfacial structure of PbI2/TiO2 is selected for subsequent investigations.
The degree of the total potential drop across the CH3NH3PbI3/TiO2 interfaces reliably indicates their photo-excited charge separation capabilities [58]. To clearly show their difference, the planar averaged electrostatic potential of the seven perovskite/TiO2 heterostructures was calculated to estimate the electronic level positions (Figure 4). The Fermi level differences between CH3NH3PbI3 and TiO2 build the driving force for the electron to transfer from the CH3NH3PbI3 to the TiO2 slab. Actually, a substantial amount of charge gather at the TiO2 side due to the abrupt potential drop near the interface. Then, the built-in electric field in the interface hampers more electron transfer across the interface, and the electronic charge transfer equilibrium is reached. It is known that the built-in electric fields originate from the surface-surface interactions, particularly for the Pb2+ movement and the CH3NH3+ orientation [58]. The incorporation of transition metals M (M = Zn2+, Y3+, Zr4+, Nb5+, Ru4+, W6+) enhances the polarization and the built-in electric field across the interfacial heterostructure. As shown in Figure 4, it is worth mentioning that the Zn- and Y-substituted interface models (x = 10Å) exhibit a substantial slope of electrostatic potential, while the others show the parameters’ electrostatic potential. The potential drop in the Zn- and Y-substituted interfaces is notably steeper than that in the other interfaces, which serves as a reservoir for electrons. Hence, a substantial number of electrons gather at the TiO2 surface, implying that the Zn2+ and Y3+ substituents are more efficient in separating the photo-generated electrons and holes.
To scrutinize the electronic property differences in various transition metal-substituted CH3NH3PbI3/TiO2, the bonding characteristics were analyzed by electron localization function (ELF), which can illustrate the type of bonding and delocalization of electron density in the interfacial system [59]. Figure 5 depicts the interfacial structures and ELF contour plots at (010) planes crossing the Pb2+ and I with color scheme for various interfacial systems. The ELF ranges from 0 to 1, where red corresponds to a full localization, blue indicates a full delocalization, and green implies the uniform electron gas. The ELF slice exhibited lesser electron localization for the transition metal substituents than Ti4+, which indicates a more covalent nature of the substituent-O interaction compared to the Ti–O interactions [60,61]. As can be seen, Figure 6a,d–g are quite similar, which explains why there is no variation in their geometry. By contrast, substitution with Zn2+ and Y3+ produced larger geometrical modification and electronic change in the CH3NH3PbI3/TiO2 interface. The ionic radius of transition metal substituents explains why there is no geometrical change with regard to the pristine and substituted systems. Compared with the Ti4+ cation (~0.6 Å) [62], the relatively large ionic radius of Y3+ (~0.9 Å) [63] and Zn2+ (~0.7 Å) [63] leads to distortions in the CH3NH3PbI3/TiO2 interface. Moreover, the binding energy and charge transfer analysis also led to the same conclusion. Interfacial bond strength varied with the ionic radius of transition metal in the substituted interfacial systems.
The optical properties, including optical reflectivity, refractive index and absorption efficient can be obtained by dielectric function [64]. Taking into account that the PCE of CH3NH3PbI3 mostly originates from the efficient use of visible light in the solar spectrum, only the calculated electronic absorption spectra of the CH3NH3PbI3/TiO2 interfaces have been investigated. The optical absorption coefficients (α) of different interfaces based on the obtained electronic structures are presented and compared in Figure 6. The absorption spectra parallel to x-axis were selected to examine the influence of transition metal substituents on the optical properties. The shapes of each absorption curve are close. As depicted in Figure 6, the CH3NH3PbI3/TiO2 interfaces have two absorption peaks—around 3.5 eV and 7.5 eV. The peak around 3.5 eV mainly comes from the conduction-to-valence band transition from I 5p or Pb 6s states to Pb 6p states [65]. On the other hand, the peak around 7.5 eV can be ascribed to the intrinsic band gap of pristine TiO2 and the electron shifting from the O 2p to Ti 3d orbitals [66]. It has also been reported that the absorption of pure TiO2 is limited to ultraviolet (UV) light and exhibits inefficient response for visible light. In case of the pristine CH3NH3PbI3/TiO2, our theoretical calculation is consistent with the experimental values and theoretical studies [67,68]. In case of the substituted CH3NH3PbI3/TiO2, the substituted interface still shows poor photoactivity in the visible-light region for solar light harvesting. However, it has an extra absorption peak in the low energy region (less than 2 eV). The incorporation of the transition metal substituents into the CH3NH3PbI3/TiO2 interface leads to an obvious red-shift effect. The distinct absorption peak at 0.3 eV in the low energy region can be attributed to the band gap near the Fermi level. The decrease in intensity of transition energies is caused by the split intra-band transitions between the impurity states, rendering the more obvious red-shift [68]. Compared with the pristine CH3NH3PbI3/TiO2 interface, the transition metal-substituted CH3NH3PbI3/TiO2 interface is expected to be more active for efficient visible-light photo-catalysis.
In order to further elucidate the charge carrier separation mechanism of the CH3NH3PbI3/TiO2 interfaces, partial density of states (PDOS) have been calculated for pure and substituted CH3NH3PbI3/TiO2 interfaces. As seen in Figure 7, the PDOS is split into contributions from CH3NH3PbI3, TiO2 and transition metal substituents. The substituent component is magnified five times for better visibility. It is already known that the band gap of TiO2 is wider than that of the CH3NH3PbI3 perovskite [25,26]. Besides, the conduction band minimum (CBM) of TiO2 is lower than that of CH3NH3PbI3. They can excite electrons from the valence band of CH3NH3PbI3 (I 5p and Pb 6s orbitals) to conduction band of CH3NH3PbI3 (Pb 6p), and then transfer to conduction band of TiO2 (Ti 3d). The difference between Pb 6p and Ti 3d decided the efficiency of charge transfer across the interface [69]. The interfacial band gaps can be tuned by n, p, and isovalent substituents using the selected transition metals. As shown in Figure 7, n substitution agents such as Nb5+ and W6+ pushed the Fermi level into the conduction band and made the system metallic. The intensity of the Fermi levels entering into the conduction band should be increased as the d states of substituents changes from 4d to 5d transition metals. By contrast, in the p substitution agents (Zn2+ and Y3+) modified interfaces, the Fermi levels shifted from valence band to conduction band, leading to obvious band gaps compared with the pristine system. For the isovalent substituted (Zr4+, Ru4+) interfaces, the PDOS shape of substituted interface is wider than that of the pristine interface, which implies that the electronic nonlocality becomes quite obvious. The delocalized transition metal d state contributes to the electron–hole pair separation in the PSC and supports carrier migration within the photo-catalysis process.
Schematic diagrams of the energy levels of various elements’ substitution TiO2 are shown in Figure 8. The vacuum level of the system was set to 0 eV for comparison. The CBM of pristine TiO2 was found to be located at −4.1 eV. It can be seen that the substituent ion effectively modified the CBM state of the TiO2. In all the cases, the CBM gradually up-shifts to the vacuum level as the substituent change from n to p. The Fermi level shifts downward with p substituent and the electron injection from perovskite to p substituted TiO2 will be hindered. On the other hand, the optimal band alignment between perovskite and n substituted TiO2 could effectively improve the charge transport and suppress charge recombination. From this point of view, Nb5+ and W6+ substituted interfaces are expected to have better device performance. Tuning energy level alignment by element substitution (M = Zn2+, Y3+, Zr4+, Nb5+, Ru4+, W6+) is confirmed to be an effective way to optimize charge transportation and thus enhance the PCE of PSC.

4. Concluding Remarks

First-principles computations were utilized to characterize the structural, electronic and optical properties of the transition metal-substituted CH3NH3PbI3/TiO2 interface. Through density functional calculations for binding energy and charge transfer of various configurations, we found that the substitution of Ti4+ at the Ti5c sites by transition metals is energetically favored. Especially, the p dopings of Zn2+ and Y3+ for Ti4+ at interfaces are the most energetically favorable among the transition metals, which lead to improved interfacial stability. Electrostatic potential investigations revealed that the potential drop in the Zn- and Y-substituted interfaces is notably steeper than that in the other interfaces, indicating that the substituents are more efficient in separating the carriers. The relatively large ionic radius of Y3+ and small ionic radius of Zn2+ lead to distortions in the ELF calculations. The calculated absorption spectra indicate that the transition metal-substituted CH3NH3PbI3/TiO2 interface retains an enhanced visible light photocatalytic ability owing to the decreased transition energies. Closer comparisons between pristine and substituted CH3NH3PbI3/TiO2 indicate that the interfacial band gaps can be tuned by n, p, and isovalent substituents using the selected transition metals. Due to their optimal band alignment, the Nb5+ and W6+ substituted interface have better device performance. Theoretical studies predict that the varied mechanisms depending on transition metal substations will exert different effects on properties of CH3NH3PbI3/TiO2 interfaces. Our calculations explain why transition metals M (M = Zn2+, Y3+, Zr4+, Nb5+, Ru4+, W6+) could enhance device performance and why it is helpful for the potential commercialization of planar heterojunction PSC.

Author Contributions

Conceptualization, X.L. and Y.G.; methodology, Y.X. and Y.G.; software, C.L.; validation, Y.X. and C.L.; resources, Y.G.; data curation, Y.X. and Y.G.; writing—original draft preparation, Y.G.; writing—review and editing, H.L., X.M., J.Z. and Z.G.; supervision, Y.N and Z.G.; project administration, Y.N.; funding acquisition, Y.N. and H.S.

Funding

This research was funded by Applied Chemistry Provincial-level Key Discipline of Henan Province.

Acknowledgments

The authors acknowledge the National Natural Science Foundation of China (Grant No. 11447150) and the Foundation of Henan Educational Committee (16B430001).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, J.; Zhou, S.; Jin, S.; Li, H.; Zhai, T. Crystal Organometal Halide Perovskites with Promising Optoelectronic Applications. J. Mater. Chem. C 2016, 4, 11–27. [Google Scholar] [CrossRef]
  2. Yang, M.; Li, Z.; Reese, M.O.; Reid, O.G.; Kim, D.H.; Siol, S.; Klein, T.R.; Yan, Y.; Berry, J.J.; van Hest, M.F.A.M.; et al. Perovskite Ink with Wide Processing Window for Scalable High-Efficiency Solar Cells. Nat. Energy 2017, 2, 17038. [Google Scholar]
  3. Chen, W.; Wu, Y.; Yue, Y.; Liu, J.; Zhang, W.; Yang, X.; Chen, H.; Bi, E.; Ashraful, I.; Grätzel, M.; et al. Efficient and Stable Large-Area Perovskite Solar Cells with Inorganic Charge Extraction Layers. Science 2015, 350, 944–948. [Google Scholar] [CrossRef] [PubMed]
  4. Heo, J.H.; Han, H.J.; Kim, D.; Ahn, T.K.; Im, S.H. Hysteresis-Less Inverted CH3NH3PbI3 Planar Perovskite Hybrid Solar Cells with 18.1% Power Conversion Efficiency. Energy Environ. Sci. 2015, 8, 1602–1608. [Google Scholar] [CrossRef]
  5. Lu, Z.; Wang, S.; Liu, H.; Feng, F.; Li, W. Improved Efficiency of Perovskite Solar Cells by the Interfacial Modification of the Active Layer. Nanomaterials 2019, 9, 204. [Google Scholar] [CrossRef] [PubMed]
  6. Mohamad Noh, M.F.; Teh, C.H.; Daik, R.; Lim, E.L.; Yap, C.C.; Ibrahim, M.A.; Ahmad Ludin, N.; Mohd Yusoff, A.R.b.; Jang, J.; Mat Teridi, M.A. The Architecture of the Electron Transport Layer for a Perovskite Solar Cell. J. Mater. Chem. C 2018, 6, 682–712. [Google Scholar] [CrossRef]
  7. Yang, G.; Tao, H.; Qin, P.; Ke, W.; Fang, G. Recent Progress in Electron Transport Layers for Efficient Perovskite Solar Cells. J. Mater. Chem. A 2016, 4, 3970–3990. [Google Scholar] [CrossRef]
  8. Ke, W.; Fang, G.; Liu, Q.; Xiong, L.; Qin, P.; Tao, H.; Wang, J.; Lei, H.; Li, B.; Wan, J.; et al. Low-Temperature Solution-Processed Tin Oxide as an Alternative Electron Transporting Layer for Efficient Perovskite Solar Cells. J. Am. Chem. Soc. 2015, 137, 6730–6733. [Google Scholar] [CrossRef]
  9. Yang, H.-Y.; Rho, W.-Y.; Lee, S.K.; Kim, S.H.; Hahn, Y.-B. TiO2 Nanoparticles/Nanotubes for Efficient Light Harvesting in Perovskite Solar Cells. Nanomaterials 2019, 9, 326. [Google Scholar] [CrossRef]
  10. Lindblad, R.; Bi, D.; Park, B.-W.; Oscarsson, J.; Gorgoi, M.; Siegbahn, H.; Odelius, M.; Johansson, E.M.J.; Rensmo, H. Electronic Structure of TiO2/CH3NH3PbI3 Perovskite Solar Cell Interfaces. J. Phys. Chem. Lett. 2014, 5, 648–653. [Google Scholar] [CrossRef]
  11. Yella, A.; Heiniger, L.-P.; Gao, P.; Nazeeruddin, M.K.; Grätzel, M. Nanocrystalline Rutile Electron Extraction Layer Enables Low-Temperature Solution Processed Perovskite Photovoltaics with 13.7% Efficiency. Nano Lett. 2014, 14, 2591–2596. [Google Scholar] [CrossRef] [PubMed]
  12. Gao, X.-X.; Ge, Q.-Q.; Xue, D.-J.; Ding, J.; Ma, J.-Y.; Chen, Y.-X.; Zhang, B.; Feng, Y.; Wan, L.-J.; Hu, J.-S. Tuning the Fermi-Level of TiO2 Mesoporous Layer by Lanthanum Doping towards Efficient Perovskite Solar Cells. Nanoscale 2016, 8, 16881–16885. [Google Scholar] [CrossRef] [PubMed]
  13. Yin, G.; Ma, J.; Jiang, H.; Li, J.; Yang, D.; Gao, F.; Zeng, J.; Liu, Z.; Liu, S.F. Correction to Enhancing Efficiency and Stability of Perovskite Solar Cells through Nb-Doping of TiO2 at Low Temperature. Acs Appl. Mater. Interfaces 2017, 9, 14545. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, B.-X.; Rao, H.-S.; Li, W.-G.; Xu, Y.-F.; Chen, H.-Y.; Kuang, D.-B.; Su, C.-Y. Achieving High-Performance Planar Perovskite Solar Cell with Nb-Doped TiO2 Compact Layer by Enhanced Electron Injection and Efficient Charge Extraction. J. Mater. Chem. A 2016, 4, 5647–5653. [Google Scholar] [CrossRef]
  15. Chen, S.-H.; Chan, S.-H.; Lin, Y.-T.; Wu, M.-C. Enhanced Power Conversion Efficiency of Perovskite Solar Cells Based on Mesoscopic Ag-Doped TiO2 Electron Transport Layer. Appl. Surf. Sci. 2019, 469, 18–26. [Google Scholar] [CrossRef]
  16. Liu, D.; Li, S.; Zhang, P.; Wang, Y.; Zhang, R.; Sarvari, H.; Wang, F.; Wu, J.; Wang, Z.; Chen, Z.D. Efficient Planar Heterojunction Perovskite Solar Cells with Li-Doped Compact TiO2 Layer. Nano Energy 2017, 31, 462–468. [Google Scholar] [CrossRef]
  17. Wu, M.-C.; Chan, S.-H.; Jao, M.-H.; Su, W.-F. Enhanced Short-Circuit Current Density of Perovskite Solar Cells Using Zn-Doped TiO2 as Electron Transport Layer. Sol. Energy Mater. Sol. Cells 2016, 157, 447–453. [Google Scholar] [CrossRef]
  18. Xu, Z.; Yin, X.; Guo, Y.; Pu, Y.; He, M. Ru-Doping in TiO2 Electron Transport Layers of Planar Heterojunction Perovskite Solar Cells for Enhanced Performance. J. Mater. Chem. C 2018, 6, 4746–4752. [Google Scholar] [CrossRef]
  19. Liu, J.; Zhang, J.; Yue, G.; Lu, X.; Hu, Z.; Zhu, Y. W-Doped TiO2 Photoanode for High Performance Perovskite Solar Cell. Electrochim. Acta 2016, 195, 143–149. [Google Scholar]
  20. Zhou, H.; Chen, Q.; Li, G.; Luo, S.; Song, T.-B.; Duan, H.-S.; Hong, Z.; You, J.; Liu, Y.; Yang, Y. Interface Engineering of Highly Efficient Perovskite Solar Cells. Science 2014, 345, 542–546. [Google Scholar] [CrossRef]
  21. Cui, Q.; Zhao, X.; Lin, H.; Yang, L.; Chen, H.; Zhang, Y.; Li, X. Improved Efficient Perovskite Solar Cells Based on Ta-Doped TiO2 Nanorod Arrays. Nanoscale 2017, 9, 18897–18907. [Google Scholar] [CrossRef] [PubMed]
  22. García-Mota, M.; Vojvodic, A.; Abild-Pedersen, F.; Nørskov, J.K. Electronic Origin of the Surface Reactivity of Transition-Metal-Doped TiO2(110). J. Phys. Chem. C 2013, 117, 460–465. [Google Scholar] [CrossRef]
  23. Pathak, S.K.; Abate, A.; Ruckdeschel, P.; Roose, B.; Gödel, K.C.; Vaynzof, Y.; Santhala, A.; Watanabe, S.-I.; Hollman, D.J.; Noel, N.; et al. Performance and Stability Enhancement of Dye-Sensitized and Perovskite Solar Cells by Al Doping of TiO2. Adv. Funct. Mater. 2014, 24, 6046–6055. [Google Scholar] [CrossRef]
  24. Xu, X.; Li, K.; Yang, Z.; Shi, J.; Li, D.; Gu, L.; Wu, Z.; Meng, Q. Methylammonium Cation Deficient Surface for Enhanced Binding Stability at TiO2/CH3NH3PbI3 Interface. Nano Res. 2017, 10, 483–490. [Google Scholar] [CrossRef]
  25. Geng, W.; Tong, C.-J.; Liu, J.; Zhu, W.; Lau, W.-M.; Liu, L.-M. Structures and Electronic Properties of Different CH3NH3PbI3/TiO2 Interface: A First-Principles Study. Sci. Rep. 2016, 6, 20131. [Google Scholar] [CrossRef] [PubMed]
  26. Ji, G.; Zheng, G.; Zhao, B.; Song, F.; Zhang, X.; Shen, K.; Yang, Y.; Xiong, Y.; Gao, X.; Cao, L.; et al. Interfacial Electronic Structures Revealed at the Rubrene/CH3NH3PbI3 Interface. Phys. Chem. Chem. Phys. 2017, 19, 6546–6553. [Google Scholar] [CrossRef] [PubMed]
  27. Feng, H.-J. Ferroelectric Polarization Driven Optical Absorption and Charge Carrier Transport in CH3NH3PbI3/TiO2-Based Photovoltaic Cells. J. Power Sources 2015, 291, 58–65. [Google Scholar] [CrossRef]
  28. Hu, J.; Ji, G.; Ma, X.; He, H.; Huang, C. Probing Interfacial Electronic Properties of Graphene/CH3NH3PbI3 Heterojunctions: A Theoretical Study. Appl. Surf. Sci. 2018, 440, 35–41. [Google Scholar] [CrossRef]
  29. Nemnes, G.A.; Goehry, C.; Mitran, T.L.; Nicolaev, A.; Ion, L.; Antohe, S.; Plugaru, N.; Manolescu, A. Band Alignment and Charge Transfer in Rutile-TiO2/CH3NH3PbI3-xClx Interfaces. Phys. Chem. Chem. Phys. 2015, 17, 30417–30423. [Google Scholar] [CrossRef]
  30. Haruyama, J.; Sodeyama, K.; Hamada, I.; Han, L.; Tateyama, Y. First-Principles Study of Electron Injection and Defects at the TiO2/CH3NH3PbI3 Interface of Perovskite Solar Cells. J. Phys. Chem. Lett. 2017, 8, 5840–5847. [Google Scholar] [CrossRef]
  31. Mosconi, E.; Grancini, G.; Roldán-Carmona, C.; Gratia, P.; Zimmermann, I.; Nazeeruddin, M.K.; De Angelis, F. Enhanced TiO2/MAPbI3 Electronic Coupling by Interface Modification with PbI2. Chem. Mater. 2016, 28, 3612–3615. [Google Scholar]
  32. Kresse, G.; Furthmüller, J. Efficiency of Ab-initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar]
  33. Blöchl, P.E. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, Y.; Yang, W. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar]
  35. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188. [Google Scholar] [CrossRef]
  36. Mosconi, E.; Umari, P.; De Angelis, F. Electronic and Optical Properties of MAPbX3 Perovskites (X = I, Br, Cl): A Unified DFT and GW Theoretical Analysis. Phys. Chem. Chem. Phys. 2016, 18, 27158–27164. [Google Scholar] [CrossRef] [PubMed]
  37. Sumita, M.; Sodeyama, K.; Jono, R.; Han, L.; Tateyama, Y. Electronic Structure of Acetonitrile Adsorbed on the Anatase TiO2 (101) Surface. Chem. Phys. Lett. 2013, 556, 225–229. [Google Scholar]
  38. Sun, R.; Wang, Z.; Shibata, N.; Ikuhara, Y. A Dislocation Core in Titanium Dioxide and Its Electronic Structure. Rsc Adv. 2015, 5, 18506–18510. [Google Scholar] [CrossRef]
  39. Fang, Y.; Cheng, D.; Niu, M.; Yi, Y.; Wu, W. Tailoring the Electronic and Optical Properties of Rutile TiO2 by (Nb+Sb, C) Codoping from DFT+U Calculations. Chem. Phys. Lett. 2013, 567, 34–38. [Google Scholar]
  40. Hou, Y.S.; Xiang, H.J.; Gong, X.G. Unveiling Magnetic Interactions of Ruthenium Trichloride via Constraining Direction of Orbital Moments: Potential Routes to Realize a Quantum Spin Liquid. Phys. Rev. B 2017, 96, 054410. [Google Scholar] [CrossRef]
  41. Gong, J.; Yang, C.; Zhang, J.; Pu, W. Origin of photocatalytic activity of W/N-codoped TiO2:H2 production and DFT calculation with GGA+U. Appl. Catal. B Environ. 2014, 152–153, 73–81. [Google Scholar] [CrossRef]
  42. Zasada, F.; Gryboś, J.; Indyka, P.; Piskorz, W.; Kaczmarczyk, J.; Sojka, Z. Surface Structure and Morphology of M[CoM′]O4 (M = Mg, Zn, Fe, Co and M′ = Ni, Al, Mn, Co) Spinel Nanocrystals—DFT+U and TEM Screening Investigations. J. Phys. Chem. C 2014, 118, 19085–19097. [Google Scholar] [CrossRef]
  43. Chen, H.-Y.T.; Tosoni, S.; Pacchioni, G. Adsorption of Ruthenium Atoms and Clusters on Anatase TiO2 and Tetragonal ZrO2(101) Surfaces: A Comparative DFT Study. J. Phys. Chem. C 2015, 119, 10856–10868. [Google Scholar] [CrossRef]
  44. Methfessel, M.; Paxton, A.T. High-Precision Sampling for Brillouin-Zone Integration in Metals. Phys. Rev. B 1989, 40, 3616–3621. [Google Scholar] [CrossRef] [PubMed]
  45. Even, J.; Pedesseau, L.; Jancu, J.-M.; Katan, C. Importance of Spin–Orbit Coupling in Hybrid Organic/Inorganic Perovskites for Photovoltaic Applications. J. Phys. Chem. Lett. 2013, 4, 2999–3005. [Google Scholar] [CrossRef]
  46. Momma, K.; Izumi, F. VESTA: A Three-Dimensional Visualization System for Electronic and Structural Analysis. J. Appl. Crystallogr. 2008, 41, 653–658. [Google Scholar] [CrossRef]
  47. Ke, W.; Fang, G.; Wang, J.; Qin, P.; Tao, H.; Lei, H.; Liu, Q.; Dai, X.; Zhao, X. Perovskite Solar Cell with an Efficient TiO2 Compact Film. Acs Appl. Mater. Interfaces 2014, 6, 15959–15965. [Google Scholar] [CrossRef] [PubMed]
  48. Mosconi, E.; Ronca, E.; De Angelis, F. First-Principles Investigation of the TiO2/Organohalide Perovskites Interface: The Role of Interfacial Chlorine. J. Phys. Chem. Lett. 2014, 5, 2619–2625. [Google Scholar] [CrossRef]
  49. Li, Z.; Wang, X.; Chi, B. First Principles Investigation of the Conversion of N2O and CO to N2 and CO2 on a Modified N + Fe/TiO2 (101) Surface. Rsc Adv. 2014, 4, 17896–17901. [Google Scholar] [CrossRef]
  50. Chen, Q.L.; Li, B.; Zheng, G.; He, K.H.; Zheng, A.S. First-Principles Calculations on Electronic Structures of Fe-Vacancy-Codoped TiO2 Anatase (101) Surface. Phys. B Condens. Matter 2011, 406, 3841–3846. [Google Scholar] [CrossRef]
  51. Ma, J.-G.; Zhang, C.-R.; Gong, J.-J.; Wu, Y.-Z.; Kou, S.-Z.; Yang, H.; Chen, Y.-H.; Liu, Z.-J.; Chen, H.-S. The Electronic Structures and Optical Properties of Alkaline-Earth Metals Doped Anatase TiO2: A Comparative Study of Screened Hybrid Functional and Generalized Gradient Approximation. Materials 2015, 8, 5508–5525. [Google Scholar] [CrossRef] [PubMed]
  52. Cromer, D.T.; Herrington, K. The Structures of Anatase and Rutile. J. Am. Chem. Soc. 1955, 77, 4708–4709. [Google Scholar] [CrossRef]
  53. Feng, J.; Xiao, B. Crystal Structures, Optical Properties, and Effective Mass Tensors of CH3NH3PbX3 (X = I and Br) Phases Predicted from HSE06. J. Phys. Chem. Lett. 2014, 5, 1278–1282. [Google Scholar] [CrossRef] [PubMed]
  54. Kawamura, Y.; Mashiyama, H.; Hasebe, K. Structural Study on Cubic-Tetragonal Transition of CH3NH3PbI3. J. Phys. Soc. Jpn. 2002, 71, 1694–1697. [Google Scholar] [CrossRef]
  55. Yang, Z.; Wang, Y.; Liu, Y. Stability and Charge Separation of Different CH3NH3SnI3/TiO2 Interface: A First-Principles Study. Appl. Surf. Sci. 2018, 441, 394–400. [Google Scholar] [CrossRef]
  56. Si, F.; Hu, W.; Tang, F.; Cheng, Y.; Xue, H. Electronic and Optical Properties of the Wurtzite- ZnO/CH3NH3PbI3 Interface: First-Principles Calculations. J. Mater. Sci. 2017, 52, 13841–13851. [Google Scholar] [CrossRef]
  57. Henkelman, G.; Arnaldsson, A.; Jónsson, H. A Fast and Robust Algorithm for Bader Decomposition of Charge Density. Comput. Mater. Sci. 2006, 36, 354–360. [Google Scholar]
  58. Feng, H.-J.; Paudel, T.R.; Tsymbal, E.Y.; Zeng, X.C. Tunable Optical Properties and Charge Separation in CH3NH3SnxPb1−xI3/TiO2-Based Planar Perovskites Cells. J. Am. Chem. Soc. 2015, 137, 8227–8236. [Google Scholar] [CrossRef]
  59. Liu, R.; Zhou, X.; Yang, F.; Yu, Y. Combination Study of DFT Calculation and Experiment for Photocatalytic Properties of S-Doped Anatase TiO2. Appl. Surf. Sci. 2014, 319, 50–59. [Google Scholar] [CrossRef]
  60. Navas, J.; Sánchez-Coronilla, A.; Aguilar, T.; Hernández, N.C.; de los Santos, D.M.; Sánchez-Márquez, J.; Zorrilla, D.; Fernández-Lorenzo, C.; Alcántara, R.; Martín-Calleja, J. Experimental and theoretical study of the electronic properties of Cu-doped anatase TiO2. Phys. Chem. Chem. Phys. 2014, 16, 3835–3845. [Google Scholar] [CrossRef]
  61. Navas, J.; Sánchez-Coronilla, A.; Gallardo, J.J.; Cruz Hernández, N.; Piñero, J.C.; Alcántara, R.; Fernández-Lorenzo, C.; De los Santos, D.M.; Aguilar, T.; Martín-Calleja, J. New Insights into Organic–Inorganic Hybrid Perovskite CH3NH3PbI3 Nanoparticles. An Experimental and Theoretical Study of Doping in Pb2+ Sites with Sn2+, Sr2+, Cd2+ and Ca2+. Nanoscale 2015, 7, 6216–6229. [Google Scholar] [CrossRef] [PubMed]
  62. Li, R.-W.; Wang, Z.-H.; Chen, X.; Sun, J.-R.; Shen, B.-G. Magnetic Properties and Colossal Magnetoresistance of the Perovskites La2/3Ca1/3Mn1−xTi xO. J. Appl. Phys. 2000, 87, 5597–5599. [Google Scholar] [CrossRef]
  63. Liu, J.; Wang, Y.; Yu, X.; Li, J. Enhanced Photoluminescence Properties of Zn2SiO4:Mn2+ Co-Activated with Y3+/Li+ under VUV Excitation. J. Lumin. 2010, 130, 2171–2174. [Google Scholar] [CrossRef]
  64. Yang, J.Y.; Liu, L.H.; Tan, J.Y. First-principles Molecular Dynamics Study on Temperature-Dependent Dielectric Function of Bulk 3C and 6H SiC in the Energy Range 3–8eV. Phys. B Condens. Matter 2014, 436, 182–187. [Google Scholar] [CrossRef]
  65. Lin, C.; Li, S.; Zhang, W.; Shao, C.; Yang, Z. Effect of Bromine Substitution on the Ion Migration and Optical Absorption in MAPbI3 Perovskite Solar Cells: The First-Principles Study. Acs Appl. Energy Mater. 2018, 1, 1374–1380. [Google Scholar] [CrossRef]
  66. Khan, M.; Xu, J.; Chen, N.; Cao, W. First Principle Calculations of the Electronic and Optical Properties of Pure and (Mo, N) Co-Doped Anatase TiO2. J. Alloy. Compd. 2012, 513, 539–545. [Google Scholar] [CrossRef]
  67. Dashora, A.; Patel, N.; Kothari, D.C.; Ahuja, B.L.; Miotello, A. Formation of an Intermediate Band in the Energy Gap of TiO2 by Cu–N-Codoping: First Principles Study and Experimental Evidence. Sol. Energy Mater. Sol. Cells 2014, 125, 120–126. [Google Scholar] [CrossRef]
  68. Jia, L.; Wu, C.; Han, S.; Yao, N.; Li, Y.; Li, Z.; Chi, B.; Pu, J.; Jian, L. Theoretical Study on the Electronic and Optical Properties of (N, Fe)-Codoped Anatase TiO2 Photocatalyst. J. Alloy. Compd. 2011, 509, 6067–6071. [Google Scholar] [CrossRef]
  69. Long, R.; Prezhdo, O.V. Dopants Control Electron–Hole Recombination at Perovskite-TiO2 Interfaces: Ab Initio Time-Domain Study. ACS Nano 2015, 9, 11143–11155. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of pristine CH3NH3PbI3/TiO2 interface models: (a) CH3NH3I/TiO2 (b) CH3NH3I/TiO2 with rotation (c) PbI2/TiO2 (d) PbI2/TiO2 with rotation.
Figure 1. Schematic illustration of pristine CH3NH3PbI3/TiO2 interface models: (a) CH3NH3I/TiO2 (b) CH3NH3I/TiO2 with rotation (c) PbI2/TiO2 (d) PbI2/TiO2 with rotation.
Nanomaterials 09 00966 g001
Figure 2. Schematic illustration of the transition metal substituents at Ti5C and Ti6C sites of the CH3NH3PbI3/TiO2 interface.
Figure 2. Schematic illustration of the transition metal substituents at Ti5C and Ti6C sites of the CH3NH3PbI3/TiO2 interface.
Nanomaterials 09 00966 g002
Figure 3. Relationship between the interfacial charge transfers and binding energies of transition metal-substituted CH3NH3PbI3/TiO2: (a) CH3NH3I/TiO2 (b) PbI2/TiO2.
Figure 3. Relationship between the interfacial charge transfers and binding energies of transition metal-substituted CH3NH3PbI3/TiO2: (a) CH3NH3I/TiO2 (b) PbI2/TiO2.
Nanomaterials 09 00966 g003
Figure 4. Planar averaged electrostatic potential across the optimized transition metal-substituted CH3NH3PbI3/TiO2.
Figure 4. Planar averaged electrostatic potential across the optimized transition metal-substituted CH3NH3PbI3/TiO2.
Nanomaterials 09 00966 g004
Figure 5. ELF of the optimized CH3NH3PbI3/TiO2 interfaces at (010) plane: (a) pristine, (b) Zn-substituted, (c) Y-substituted, (d) Zr-substituted, (e) Nb-substituted, (f) Ru-substituted, (g) W-substituted.
Figure 5. ELF of the optimized CH3NH3PbI3/TiO2 interfaces at (010) plane: (a) pristine, (b) Zn-substituted, (c) Y-substituted, (d) Zr-substituted, (e) Nb-substituted, (f) Ru-substituted, (g) W-substituted.
Nanomaterials 09 00966 g005
Figure 6. Comparison of the optical absorption of the transition metal-substituted CH3NH3PbI3/TiO2 interfaces.
Figure 6. Comparison of the optical absorption of the transition metal-substituted CH3NH3PbI3/TiO2 interfaces.
Nanomaterials 09 00966 g006
Figure 7. DOS of pristine and transition metal-substituted CH3NH3PbI3/TiO2 interface.
Figure 7. DOS of pristine and transition metal-substituted CH3NH3PbI3/TiO2 interface.
Nanomaterials 09 00966 g007
Figure 8. Schematic energy level diagram of element substituted TiO2.
Figure 8. Schematic energy level diagram of element substituted TiO2.
Nanomaterials 09 00966 g008
Table 1. Calculated lattice parameters of TiO2 and MAPbI3 and deviation between experiment and simulation.
Table 1. Calculated lattice parameters of TiO2 and MAPbI3 and deviation between experiment and simulation.
Anatase TiO2Tetragonal -MAPbI3
acac
Experimental3.7859.5148.8012.685
Calculated3.799.538.8013.05
Deviation0.13%0.17%-2.8%
Table 2. Interfacial binding energy (in eV) and lattice mismatch of the CH3NH3PbI3/TiO2 interfaces.
Table 2. Interfacial binding energy (in eV) and lattice mismatch of the CH3NH3PbI3/TiO2 interfaces.
CH3NH3I/TiO2PbI2/TiO2CH3NH3I/TiO2
With Rotation
PbI2/TiO2
With Rotation
Binding energy2.162.070.001.44
Lattice mismatch−12.0%−12.0%−13.8%−13.8%
Charge transfer−0.29−0.28−0.16−0.16
Table 3. Total energy (in eV) of CH3NH3PbI3/TiO2 with substitution of one Nb5+ for Ti4+.
Table 3. Total energy (in eV) of CH3NH3PbI3/TiO2 with substitution of one Nb5+ for Ti4+.
PositionSurface
(Ti5c)
Sub-Surface
(Ti6c)
Inner-Surface
(Third Ti Layer)
Inner-Surface
(Fourth Ti Layer)
Total energy−1401.71−1401.94−1401.52−1401.72
Table 4. Interfacial binding energy (in eV) of the Nb-substituted and pristine CH3NH3PbI3/TiO2 interfaces.
Table 4. Interfacial binding energy (in eV) of the Nb-substituted and pristine CH3NH3PbI3/TiO2 interfaces.
Ti5c SiteTi6c SiteCH3NH3I/TiO2PbI2/TiO2
CH3NH3I/TiO2PbI2/TiO2CH3NH3I/TiO2PbI2/TiO2
Binding energy3.593.052.772.540.090.00
Charge transfer−0.44−0.28−0.26−0.15−0.29−0.28

Share and Cite

MDPI and ACS Style

Guo, Y.; Xue, Y.; Li, X.; Li, C.; Song, H.; Niu, Y.; Liu, H.; Mai, X.; Zhang, J.; Guo, Z. Effects of Transition Metal Substituents on Interfacial and Electronic Structure of CH3NH3PbI3/TiO2 Interface: A First-Principles Comparative Study. Nanomaterials 2019, 9, 966. https://doi.org/10.3390/nano9070966

AMA Style

Guo Y, Xue Y, Li X, Li C, Song H, Niu Y, Liu H, Mai X, Zhang J, Guo Z. Effects of Transition Metal Substituents on Interfacial and Electronic Structure of CH3NH3PbI3/TiO2 Interface: A First-Principles Comparative Study. Nanomaterials. 2019; 9(7):966. https://doi.org/10.3390/nano9070966

Chicago/Turabian Style

Guo, Yao, Yuanbin Xue, Xianchang Li, Chengbo Li, Haixiang Song, Yongsheng Niu, Hu Liu, Xianmin Mai, Jiaoxia Zhang, and Zhanhu Guo. 2019. "Effects of Transition Metal Substituents on Interfacial and Electronic Structure of CH3NH3PbI3/TiO2 Interface: A First-Principles Comparative Study" Nanomaterials 9, no. 7: 966. https://doi.org/10.3390/nano9070966

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