Haiming Huang*ab,
Mingquan Dingab,
Yu Zhangab,
Shuai Zhangab,
Yiyun Linga,
Weiliang Wang*c and
Shaolin Zhang*ab
aSolid State Physics & Material Research Laboratory, School of Physics and Materials Science, Guangzhou University, Guangzhou 510006, China. E-mail: huanghm@gzhu.edu.cn; slzhang@gzhu.edu.cn
bResearch Center for Advanced Information Materials (CAIM), Huangpu Research and Graduate School of Guangzhou University, Guangzhou 510555, China
cSchool of Physics, Guangdong Province Key Laboratory of Display Material and Technology, Sun Yat-sen University, Guangzhou 510275, China. E-mail: wangwl2@mail.sysu.edu.cn
First published on 18th May 2023
Hybrid organic switch-inorganic semiconductor systems have important applications in both photo-responsive intelligent surfaces and microfluidic devices. In this context, herein, we performed first-principles calculations to investigate a series of organic switches of trans/cis-azobenzene fluoride and pristine/oxidized trimethoxysilane adsorbed on low-index anatase slabs. The trends in the surface–adsorbate interplay were examined in terms of the electronic structures and potential distributions. Consequently, it was found that the cis-azobenzene fluoride (oxidized trimethoxysilane)-terminated anatase surface attains a lower ionization potential than the trans-azobenzene fluoride (pristine trimethoxysilane)-terminated anatase surface due to its smaller induced (larger intrinsic) dipole moment, whose direction points inwards (outwards) from the substrate, which originates from the electron charge redistribution at the interface (polarity of attached hydroxyl groups). By combining the induced polar interaction analysis and the experimental measurements in the literature, we demonstrate that the ionization potential is an important predictor of the surface wetting properties of adsorbed systems. The anisotropic absorbance spectra of anatase grafted with azobenzene fluoride and trimethoxysilane are also related to the photoisomerization and oxidization process under UV irradiation, respectively.
In this case, a good understanding of hybrid systems is crucial to quantify the interplay between the organic switches and inorganic substrates during the photochemical process. To explore the optoelectronic properties of these hybrid systems, numerous studies have carried out, including density functional theory (DFT)/time-dependent density functional theory (TDDFT) investigation of a series of photochromic derivatives of trans/cis-AZB,33 open/closed-SP34 and open/closed-DTE33,35 adsorbed onto anatase and rutile slabs. To explain the switchable wettability of TiO2 nanoparticles modified with trimethoxysilane under UV illumination, a series of experiments was carried out to confirm the oxidization process for the degradation of trimethoxysilane.24,25 However, complementary computational efforts to model the surface of these systems and interface properties are lacking.
Surface potential is one of the basic physical characteristics of a material surface, which is related to its surface properties (through the Fermi energy), given that it can dictate the direction of charge transfer, and thereby significantly affect the stability and surface chemistry of surfaces with adsorbates.36–38 Particularly, the work function (ionization potential) of the metallic (semiconducting or insulating) surface is one of the most important technological parameters of the surface potential39,40 in developing new field-emitter cathodes or photo-responsive superhydrophobic surfaces for applications in light-emitting diodes and smart fluid control. By using the surface polarizability approach and performing contact angle measurements, the surface polarizability, which has been observed to exhibit a significant inverse relationship with respect to the ionization potential,41–43 was found to be closely related to the surface free energy and wettability.44 By combining first-principle calculation with contact angle measurement, the work function was demonstrated to be a critical quantity in understanding the wettability of silane coatings.45 Based on first-principles density functional theory, low-dimensional materials with a low work function have been demonstrated to have a considerable impact on the power efficiency for field-emission applications.46 However, a limited number of studies has been conducted on the surface potentials of joint systems that incorporate organic switches in inorganic surfaces. In this case, how the functional organic switch configuration, low-index semiconductor surface and surface–adsorbate interplay influence the electronic properties and surface potential barriers of the joint system remains an unsolved issue. Furthermore, the difference in surface potentials between a series of organic switches grafted on various inorganic semiconductor surfaces still needs to be studied in depth.
Herein, we used DFT to investigate the electronic properties and surface potentials of hybrid organic/inorganic systems with different organic adsorbates and various low-index TiO2 surfaces. We show that the change in the ionization potential during the photoisomerization and oxidization processes of the organic switch configuration is correlated with the changes in the surface dipole moments. We demonstrate that the ionization potentials play an important role in determining the potential barriers, and consequently justify their surface-wetting properties. This work is organized as follows: the Computational methods section provides the physical model and the computational methodology. The results and their interpretations are presented in the Results and discussion section. Finally, the Conclusions section provides a general summary of this work.
We carried out DFT calculations with the Perdew–Burke–Ernzerhof (PBE)53 form of the generalized gradient approximation (GGA) for the exchange-correlation functional to optimize the atomic structures and calculate the electronic structures, as implemented in the Vienna ab initio simulation package (VASP).54,55 The ion–electron interactions were treated using the projector augmented wave (PAW) method56 with a cutoff energy of 400 eV. The two-dimensional joint system was modeled by using the supercell approximation with at least a 25 Å vacuum space along the direction perpendicular to the surface, which was taken to be the z-axis, to eliminate interactions between periodic images. From the bulk anatase cell, a four-layer 4 × 1 and 1 × 4 slab was cut in the (100) and (101) directions, presenting dimensions of 11.36 × 9.87 Å2 and 10.57 × 11.36 Å2 in the x–y plane, respectively, and the bottom layer along the z direction was fixed. All the selected atoms were fully relaxed using the conjugate gradient method until the magnitude of the force on each atom became less than 0.01 eV Å−1 in the optimized structures and the total energies were converged within 10−5 eV per cell. For the structural optimization (static calculations), the Brillouin zone was sampled using a set of 2 × 2 × 1 (4 × 4 × 1) Monkhorst–Pack k points.57 The Gaussian smearing scheme was applied with a smearing width of 0.14 eV. The adsorption coverage in our calculation corresponded to one organic adsorbate per TiO2 4 × 1 (1 × 4) supercell for the (100) [(101)] surface, i.e., ∼8.9 × 1013 cm−2 (∼8.3 × 1013 cm−2) for (100) [(101)] surface. The coverage we considered is in the reasonable range reported in previous experimental and theoretical studies.48,49,58
The ionization potential (work function) for a semiconductor (metal) was derived from the energy difference between the vacuum level and the valence band maximum (Fermi energy). Generally, the Fermi level is located between the conduction band minimum and the valence band maximum depending on the hybrid and surface states. Given that the adsorption is on only one side of the hybrid organic/inorganic surface, in the calculation of surface potentials, we put two joint systems with mirror symmetry in the supercell to create two vacuum regions with constant vacuum potentials to obtain two independent vacuum levels of the two sides of the surface. It is noteworthy that our computational approach was validated by performing sets of work function calculations on silane-modified substrates45 and graphene edges.37
Structure (as labeled in Fig. 1 and S1) | Total energy (eV) | Adsorption energy (eV) |
---|---|---|
trans-FAZB/(100) | −834.80 | −2.68 |
cis-FAZB/(100) | −834.32 | −2.59 |
trans-FAZB/(101)-I | −826.50 | −2.63 |
cis-FAZB/(101)-I | −825.94 | −2.54 |
trans-FAZB/(101)-II | −835.04 | −2.14 |
cis-FAZB/(101)-II | −834.68 | −2.15 |
PFOS-F/(100) | −839.55 | −3.05 |
PFOS-OH/(100) | −845.91 | −3.09 |
PFOS-F/(101)-I | −832.01 | −3.37 |
PFOS-OH/(101)-I | −838.37 | −3.68 |
PFOS-F/(101)-II | −840.90 | −2.73 |
PFOS-OH/(101)-II | −849.15 | −3.16 |
NO3/(100) | −654.78 | −1.08 |
NO3/(101)-I | −642.26 | −2.00 |
NO3/(101)-II | −652.64 | −1.07 |
Given that all the surface properties of the organic/inorganic hybrid systems were derived by truncation of the bulk systems and modification of the initial surfaces, in Fig. S2† and 2 we reported the electronic structures of the TiO2 bulk and three different pristine anatase surfaces. An indirect band gap of 2.13 eV between the Γ and X points was found for bulk anatase TiO2, which is significantly smaller than the experimental band gap value of 3.2 eV (ref. 60) but in agreement with the theoretical studies (2.0 ∼ 2.37 eV).61,62 Our PBE-computed band gaps for the unbound (100), (101)-I and (101)-II anatase slabs are 2.11, 2.18 and 2.46 eV, respectively, in agreement with the previous studies on TiO2 slab systems (1.75 ∼ 2.5 eV).33 Fig. 2(d) shows the typical potential energy distribution across the unbound TiO2 surface. The vacuum potential at a remote location far away from the surface in the z direction is an asymptotic constant. Given that the valence band maximum was set to zero in our calculation, the vacuum level, which is determined from the electrostatic potential in the vacuum region far enough away from the slab surface, is equal to the ionization potential. For each unbound surface, the difference in ionization potentials between the fixed bottom side and the relaxed top side is influenced by surface relaxation. It is clear that the unbound (101)-II [(101)-I] surface has the highest (lowest) ionization potential.
Here, we consider the impact of adsorption on the electronic structure of the hybrid systems as given by the energy band. The band structures, as depicted in Fig. 3 for the FAZB/TiO2 systems and Fig. 4 for the PFOS/TiO2 systems, show the contribution of the organic adsorbates. For a given TiO2 substrate, all the joint systems are semiconductors with band gaps that are smaller than that of the unbound surface. A comparison of the energy gaps between the hybrid and unbound systems is presented in Table 2, where we notice that for each type of TiO2 substrate, the band gap of the hybrid system with cis-FAZB adsorption is smaller than that with trans-FAZB adsorption, while the band gaps of the hybrid systems with PFOS-F and PFOS-OH adsorption show small differences.
Structure (as labeled in Fig. 1 and S1) | Energy gap (eV) | Ionization potential (eV) | |
---|---|---|---|
Modified side (relaxed side) | Clean side (fixed side) | ||
(100) | 2.11 | 5.86 | 4.62 |
(101)-I | 2.18 | 5.76 | 3.96 |
(101)-II | 2.46 | 6.23 | 5.47 |
trans-FAZB/(100) | 1.77 | 5.17 | 4.90 |
cis-FAZB/(100) | 1.62 | 4.61 | 4.88 |
trans-FAZB/(101)-I | 1.62 | 5.48 | 3.92 |
cis-FAZB/(101)-I | 1.54 | 5.00 | 3.85 |
trans-FAZB/(101)-II | 1.84 | 5.07 | 5.69 |
cis-FAZB/(101)-II | 1.54 | 4.44 | 5.68 |
PFOS-F/(100) | 1.78 | 6.24 | 4.61 |
PFOS-OH/(100) | 1.79 | 5.79 | 4.69 |
PFOS-F/(101)-I | 1.86 | 6.75 | 3.85 |
PFOS-OH/(101)-I | 1.84 | 6.02 | 3.88 |
PFOS-F/(101)-II | 2.16 | 6.05 | 5.55 |
PFOS-OH/(101)-II | 2.27 | 5.46 | 6.08 |
Given that both the FAZB/TiO2 and PFOS/TiO2 hybrid systems are semiconductors, we explored their ionization potential to quantify the energy barrier to free space that prevents an electron at the valence band maximum from escaping the solid. Fig. 5(a) and (c) show the potential energy distribution of FAZB and PFOS adsorbed on different anatase surfaces, respectively. We plotted the electrostatic potential energy along a line that crosses the adsorbates and is perpendicular to the slab. It is clear that the potential energy in the vicinity of the surface increases gradually as the distance increases for both the adsorbed side and the clean side [Fig. 5(a and c), respectively], but the variation in the potential energy is not monotonous in the former case due to the existence of adsorbates. We found that the vacuum level of the adsorbed side is different from that of the clean side. The difference in the vacuum levels between the adsorbed side and the clean side is determined by the adsorbate type. It is obvious that the vacuum levels of the clean sides are independent of the type of organic adsorbate on the other side and similar to that of the unbound surfaces without relaxation. To further investigate the potential barriers of the TiO2 surface with different organic switch modifications, we plotted the potential energy curves along the forward path, which is the line perpendicular to the surface, with the origin at the outermost atoms, as shown in Fig. 5(b and d), respectively. The corresponding ionization potentials obtained from the vacuum levels in Fig. 5(b and d) are summarized in Table 2. We noted that the ionization potentials of the FAZB/TiO2 hybrid systems follow the order of trans-FAZB/(101)-I > trans-FAZB/(100) > trans-FAZB/(101)-II > cis-FAZB/(101)-I > cis-FAZB/(100) > cis-FAZB/(101)-II. We found that the adsorption on a semiconducting substrate with each given FAZB configuration changes the relative order of the ionization potential between the different TiO2 surfaces compared to that without adsorption. For example, the relative order of the ionization potential for the (100), (101)-I and (101)-II hybrid surfaces grafted with trans-FAZB is different from that for the (100), (101)-I and (101)-II unbound surfaces. Clearly, regardless of the type of anatase surface, the ionization potential of the TiO2 surface grafted with trans-FAZB is always larger than that with cis-FAZB. Alternatively, the ionization potential of the TiO2 surface modified with various short carbon chain silanes follows the order of PFOS-F/(101)-I > PFOS-F/(100) > PFOS-F/(101)-II > PFOS-OH/(101)-I > PFOS-OH/(100) > PFOS-OH/(101)-II. It is clear that ionization potentials of the different types of TiO2 surfaces modified with a given silane exhibit the same relative order as that of the TiO2 surfaces modified with a given FAZB. The PFOS-F-grafted TiO2 surface also shows a larger ionization potential than the PFOS-OH-grafted TiO2 surface. Notably, the hierarchy of the ionization potential of the hybrid systems predicted by our numerical calculations was found to follow the same trend of hydrophobicity in experimental measurements [105° (23°) contact angle for the trans-state (cis state) FAZB-grafted silica-encapsulated nanoparticles23 and 165° (<10°) contact angle for PFOS-F (PFOS-OH)-grafted TiO2 nanoparticles25], which can be explained by the induced polar interaction model at the liquid/solid interfaces.44 When a polar liquid (e.g., water) is in contact with a polarizable solid surface, the induced solid/liquid polar interaction, and thereby the surface hydrophilicity increases with the solid surface polarizability monotonously, which is inversely related to the ionization potential.41–43 Thus, we can expect that the equilibrium contact angle for contacting a polar liquid will increase with an increase in the ionization potential of solid surface, i.e., the surface hydrophobicity of trans-FAZB (PFOS-F)-grafted TiO2 is larger than that of cis-FAZB (PFOS-OH)-grafted TiO2.
To understand how the organic adsorbates affect the surface potential, we considered in detail the redistributions of the electron charges between the inorganic substrate and the organic adsorbate, as shown in Fig. 6 and 7. A common feature is that the charge redistributions are locally confined in the vicinity of the interface between the TiO2 surface and the adsorbates. In the case of the FAZB/TiO2 (PFOS/TiO2) hybrid system, the positive charges accumulate on the top layer of the TiO2 substrate [indicated in Fig. 6 (Fig. 7)], while negative charges accumulate at the binding site of FAZB (PFOS) [indicated in Fig. 6 (Fig. 7)]. However, the end-to-end distance and the linear distribution of transferred electron charge for the planar rod-like trans-FAZB is larger than that of nonplanar cis-FAZB, suggesting that the trans-FAZB receives more electrons from the interface and induces a larger dipole moment whose direction points inwards from the TiO2 substrate. It had been reported that the adsorbate on the substrate reduces (enhances) the electron binding capability, i.e., work function or ionization potential, by forming a plane of dipoles, whose direction points outwards (inwards) from the substrate.37 Therefore, compared to the case of the cis-FAZB-grafted TiO2 surface, the higher ionization potential for the trans-FAZB-grafted TiO2 surface can be attributed to the larger electron charge transfer caused by chemical functionalization and structural relaxation effect. Alternatively, the relative order of ionization potential between the PFOS-F- and PFOS-OH-grafted TiO2 surfaces cannot be explained by the effect of the induced dipole moment derived from charge transfer only. As shown in Fig. 7, for a given TiO2 surface, the linear distribution of transferred electron charge of PFOS with hydroxyl substitution is almost the same as that of pristine PFOS. It is widely known that the hydroxyl group is a polar chemical group with an intrinsic dipole moment pointing from oxygen to hydrogen. Fig. 7(d) shows the dipole moment contributions of the PFOS-F and PFOS-OH adsorbates. The contribution of the OH dipole moment to the potential barrier is negative given that its direction points outwards to the anatase slab [Fig. 7(d)]. This explains why the OH-substituted PFOS exhibits exceptionally low ionization potentials. The effect of the intrinsic dipole moments of the functional groups was also predicted with an increase or decrease in the potential barriers of several surfaces.46,63
Also, we calculated the UV-vis absorbance spectra of the hybrid organic/inorganic systems to probe the grafting-induced changes in their optical properties. Consequently, a significant anisotropic absorbance was found for the bare and hybrid surfaces. The first absorption bands in Fig. 8 are 3.02, 2.74 and 3.39 eV for the (100), (101)-I and (101)-II unbound TiO2 surfaces, respectively. Our computed first absorption band of the (101)-I unbound surface is consistent with other theoretical calculations (2.56 eV)34 but is underestimated compared with the experimental measurements, which is located between 3.4 and 4.9 eV performed on TiO2 films with different thicknesses.64 To obtain the exact optical properties in the UV-vis range, the scissors operation65 of 0.66 eV, which corresponds to the difference between the experimental measurement of the thinnest TiO2 films (∼15 nm) considered in a previous study and our theoretical result, was carried out in an optical absorbance of the hybrid system. As shown in Fig. 9(a), we found that the absorbance spectra of the TiO2 surface grafted with trans-FAZB are slightly different from that with cis-FAZB, except for the case of TiO2 surface grafted with trans-FAZB in the z polarization direction, which exhibits an additional absorption peak at about 3.2 eV, corresponding to the conformational switch of azobenzene-containing units from the apparent wetting state (i.e., trans isomer) to the fully wetting state (i.e., cis isomer) under UV irradiation.23 In addition, the absorbance spectra of the (101)-II TiO2 surface grafted with PFOS-F are also slightly different from that with PFOS-OH [Fig. 9(b)], while the absorbance spectra of PFOS-F- and PFOS-OH-grafted TiO2 surfaces are similar for the (100) and (101)-I surfaces. The absorption peak at about 3.6 eV for the PFOS-grafted TiO2 surface can be assigned to the oxidization of PFOS upon UV illumination.25
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3ra00537b |
This journal is © The Royal Society of Chemistry 2023 |