Mengjiao
Wang
*a,
Silvio
Osella
b,
Rosaria
Brescia
c,
Zheming
Liu
d,
Jaime
Gallego
a,
Mattia
Cattelan
e,
Matteo
Crisci
a,
Stefano
Agnoli
e and
Teresa
Gatti
*af
aInstitute of Physical Chemistry and Center for Materials Research (LaMa), Justus Liebig University, 35392 Giessen, Germany. E-mail: mengjiao.wang@phys.chemie.uni-giessen.de
bChemical and Biological Systems Simulation Lab, Centre of New Technologies, University of Warsaw, 02097 Warsaw, Poland
cElectron Microscopy Facility, Istituto Italiano di Tecnologia, Via Morego, 30, 16163 Genova, Italy
dNanochemistry Department, Istituto Italiano di Tecnologia, 16163 Genova, Italy
eDepartment of Chemical Sciences, University of Padova, 35131 Padova, Italy
fDepartment of Applied Science and Technology, Politecnico di Torino, 10129 Torino, Italy. E-mail: teresa.gatti@polito.it
First published on 25th November 2022
As a semiconductor used for the photocatalytic hydrogen evolution reaction (HER), BiOBr has received intensive attention in recent years. However, the high recombination of photoexcited charge carriers results in poor photocatalytic efficiency. The combination with other photoactive semiconductors might represent a valuable approach to deal with the intrinsic limitations of the material. Given that BiOBr has a 2D structure, we propose a simple liquid-phase exfoliation method to peel BiOBr microspheres into few-layer nanosheets. By tuning the weight ratio between the precursors, we prepare a series of 2D MoS2/BiOBr van der Waals (vdW) heterojunctions and study their behaviour as (photo)electrocatalysts for the HER, finding dramatic differences as a function of weight composition. Moreover, we found that pristine 2D BiOBr and the heterojunctions, with the exception of the 1% MoS2/BiOBr composition, undergo photocorrosion, with BiOBr being reduced to metallic Bi. These findings provide useful guidelines to design novel 2D material-based (photo)electrocatalysts for the production of sustainable fuels.
BiOBr is emerging as an interesting photocatalyst because of its unique internal electric field and band structure, which facilitate the separation and mobility of the charge carriers. Moreover, the layered structure, kept together by van der Waals (vdW) interactions, makes BiOBr suitable for exfoliation into 2D nanosheets, with a high surface area available for catalytic reactions. There are already a number of published works studying the photocatalytic processes promoted by 2D BiOBr, such as the photodegradation of pollutants,13–15 photoreduction of CO2,16 photocatalytic nitrogen fixation17 and photocatalytic water splitting.18 However, like most of the photocatalysts that are still at the laboratory stage development, 2D BiOBr is in its research infancy as well. The application of an additional voltage to drive the catalytic reaction increases the efficiency of photoreactions, but the stability of BiOBr in the PEC environment can be strongly compromised. Originally, 2D BiOBr was used as a photoanode to produce O2 during water splitting, mainly because the energy position of its valence band (VB) energy is suitable for water oxidation, while the conduction band (CB) is not properly positioned for the reduction of H+/H2.19 However, recently a growing number of reports described BiOBr-based photocathodes as photoelectrocatalysts for the hydrogen evolution reaction (HER). In this context, different methods have been applied to tune the band structure of BiOBr to increase the PEC efficiency. For example, Shi et al. synthesized facet-dependent BiOBr and found that samples with dominating (001) facets have a band structure that is appropriate for proton reduction and performs excellent HER.18 Wang et al. applied a chemical vapor deposition method to grow 2D BiOBr for the PEC HER.20 Despite its promising catalytic activity, the photoreduction of BiOBr to metallic Bi remains a bottleneck to overcome to maintain the catalyst stability.19 Different strategies have been applied to enhance both the efficiency and stability of BiOBr. Among them, the generation of heterojunctions (HJs) with other materials was particularly effective and various examples are present in the literature. Theoretically, a HJ establishing a fine-tuned internal electric field is helpful in separating and transferring the photogenerated electrons or holes from the original photoelectrocatalyst, in order to reduce the charge recombination and hinder the photocorrosion reaction. Li et al. reported an oxygen vacancy-BiOBr/Cu2−xS type I HJ with a high hydrogen evolution rate, in which the divalent defects in the HJ are able to create interfacial polarization, and this polarization has a synergetic effect with the internal electric field.21 This work indeed inspired us to fabricate 2D BiOBr HJs with other materials, to further explore the potential of this approach to boost the PEC HER efficiency and stability in this catalyst.
The design of vdW HJs has been pioneered by combining different 2D layered materials through facile synthesis methods.22 The construction of vdW bonds has already proved to be a promising technique for boosting the catalytic properties of 2D materials.23 MoS2 is a popular 2D layered material used for building vdW HJs with other layered materials.24 These vdW HJs, such as black phosphorus/MoS2,25 graphene/MoS2,26 WSe2/MoS227 and MoS2/MoSe2,28 can provide access to variable bandgaps, fast charge carrier transport or interfacial coupling effects. There are several works reporting on MoS2/BiOBr HJs, and a variety of synthesis methods such as mechanical ball milling,29 chemical deposition30 and wet chemical co-precipitation methods31 have been applied to fabricate similar nanoarchitectures with different morphologies. These materials have been mainly applied in the photocatalytic degradation of pollutants, but the analyses of the band structure and the type of HJ present in these systems show significant discrepancies, thus the photocatalytic mechanism is still unclear. Moreover, to the best of our knowledge, a 2D MoS2/BiOBr vdW HJ has not been reported yet.
Therefore, in this work, we used a solvothermal route to prepare BiOBr microspheres and liquid-phase exfoliation (LPE) to pretreat a bulk MoS2 powder. Then, we employed another LPE process on the mixture of BiOBr microspheres and pre-treated MoS2 to obtain 2D BiOBr/MoS2 vdW HJs (Scheme 1). The products, with different MoS2/BiOBr ratios, were tested in the PEC HER. Surprisingly, only the 1% (w/w) MoS2/BiOBr HJ has a remarkable catalytic activity among all investigated samples. In this system, during PEC, BiOBr absorbs UV radiation and generates separated electrons and holes, while MoS2 accepts the photogenerated electrons and performs the HER, while avoiding BiOBr photocorrosion at the same time. Our findings provide a novel method for synthesizing MoS2/BiOBr vdW HJs and a new strategy to modify the efficiency and stability of 2D material-based photoelectrocatalysts.
![]() | ||
Scheme 1 Sketch of the process employed in this work to synthesize 2D MoS2/BiOBr HJs. Solvothermal and LPE methods have been applied to yield inks of the layered binary nanostructure in NMP. |
The powder X-ray diffraction (P-XRD) patterns of 2D BiOBr and 2D MoS2 reveal the presence of the tetragonal BiOBr phase (ICSD number: 61225) and hexagonal MoS2 phase (ICSD number: 95569) without any impurities (Fig. 2a). Except for the 1% MoS2/BiOBr sample, which has a low MoS2 content and displays only the BiOBr structure, the P-XRD patterns of the MoS2/BiOBr HJs feature both phases with an increasing peak intensity ratio of MoS2 (002) (located at 14.4°) to BiOBr (102) (located at 31.7°) when the weight ratio of MoS2 increases. This is because the XRD peak intensity at 14.4° is related to the extended (001) planes in MoS2 2D sheets, which increases because of the increasing amount of the MoS2 component. It is worth mentioning that the characteristic peak appeared at 10.4°, which is corresponding to the (001) direction of the tetragonal BiOBr phase, is not clearly detected in the BiOBr precursor or LPE 2D BiOBr, while it is detected in the four HJ samples. This can be due to the fact that the combination with MoS2 enhances the stacking of BiOBr layers along the (001) direction. Raman spectra were also recorded to identify possible vdW interactions between BiOBr and MoS2. The characteristic peaks of both components were distinguished in all HJs (Fig. 2b). Specifically, the peaks at 113.2 and 160.6 cm−1 correspond to the A1g and Eg mode of internal Bi–Br stretching, respectively, while the peaks at around 381.3 and 405 cm−1 correspond to the in-plane (E2g) mode of Mo–S and out-of-plane (A1g) mode of S–S.33 Clearly, the intensity of BiOBr characteristic peaks is reduced dramatically with the progressive increase in the MoS2 content, indicating a strong interaction between the two components. As for the two characteristic peaks of MoS2, a shift to higher wavenumbers for both peaks is observed for the HJs compared to 2D MoS2, further pointing out a possible vdW connection between MoS2 and BiOBr. To analyse the characteristic peaks of MoS2 better, we show a magnification of the region between 360 and 430 cm−1 in Fig. 2c. It is reported that the distance between the two fingerprint peaks (E2g and Ag) of MoS2 has a strong correlation with the number of layers.34–36 As for the bare LPE MoS2, we find a wavenumber difference of 24.3 cm−1, indicating that the average number of layers in the sample is around 3. On the other hand, for all the HJs presented here, an observed wavenumber difference of over 25 cm−1 between the two representative peaks probably indicates an increased number of MoS2 layers. However, this accumulation of layers can be attributed both to MoS2–MoS2 layer restacking and MoS2–BiOBr restacking, with the second effect being predominant at a lower MoS2 concentration in the HJs due to the higher probability of the MoS2 sheet to interact with a BiOBr layer. This is also indirectly proven by the wavenumber shift of both peaks. It is known that the Raman peak energy is strongly related to the polarization of the constituent atoms in the phonon vibration, with bonds involving softer atoms requiring lower energies to vibrate. Since oxygen is largely present in BiOBr (the O/Bi atomic ratio is more than 1, see Table S1†), it is possible that oxygen atoms located at the interface of the HJ interact with sulphur vacancies on MoS2, thus causing an average decrease in polarizability in Mo–S(O) and S–S(O) bonds (due to the harder character of O compared to S), that drives the frequency shift to higher wavenumbers, as it was also observed for in situ oxygen-doped MoS2 monolayers.37
UV-visible absorption spectra (Fig. 3a and b) allow us to estimate the light-harvesting features and, by applying the Tauc method, also the band gaps of the LPE HJs. As expected, both 2D BiOBr (3.15 eV) and 2D MoS2 (1.75 eV) show wider band gaps compared to their corresponding bulk material precursors.38,39 In addition, the band gap of the vdW HJs progressively narrows with the addition of MoS2, from 3.01 eV to 2.25 eV (see details in Table S2†).40,41 Photoluminescence (PL) spectra were also measured at an excitation wavelength of 375 nm, which was determined to be the wavelength providing higher PL intensity by recording the excitation spectrum of bare 2D BiOBr (Fig. S7†). As shown in Fig. 3c, both 2D BiOBr and the HJs feature a broad emission, with a peak located at around 451.7 nm for pure 2D BiOBr and 477.7 nm for the HJs. Considering that a large Stokes shift (∼0.4 eV) is present between the absorption peak (3.15 eV) and the PL emission (2.74 eV), it is safe to assume that this PL generates from an exciton emission related to the defects and surface oxygen vacancy of 2D BiOBr.42 Besides, the absence of the band edge emission, which should be at around 400 nm, suggests weak recombination of photogenerated charges inside 2D BiOBr. However, the intensity of PL decreases with the increasing content of 2D MoS2, which is an important signature that the optical recombination path in 2D BiOBr is strongly influenced by the presence of nearby MoS2 nanosheets.
To gain further insight into the electronic properties of the HJs constructed by different ratios of MoS2 and BiOBr, density functional theory (DFT) calculations were performed on three simplified HJ models by tuning the layer ratios of MoS2 and BiOBr, namely 1 layer of MoS2 and 2 layers of BiOBr (1-MoS2/2-BiOBr), 1 layer of MoS2 and 1 layer of BiOBr (1-MoS2/1-BiOBr), and 2 layers of MoS2 and 1 layer of BiOBr (2-MoS2/1-BiOBr). The optimized structures of the three different interfaces are reported in Fig. 4a, S8a and S8c.† As can be clearly identified in all of them, the BiOBr sheet always exposes the oxygen-rich face towards the interface with MoS2 and not towards the bromine-rich one. This further corroborates the previously discussed observations on Raman features of MoS2 in the HJs. Our calculations further suggest that the band structure and charge transfer are influenced by different layer combinations. From the total and partial density of state (DOS) analysis of different interfaces we can observe a prevalent MoS2 character for the valence band in all three models, while for the valence band (VB) differences arise depending on the layer ratio difference (Fig. 4b, Fig. S8b and S8d†). In particular, only for the 1-MoS2/1-BiOBr HJ the VB is fully localized on the BiOBr, while for the other two interfaces mixed contributions arising from both MoS2 and BiOBr are present, confirming the strong charge recombination effect that the presence of MoS2 has on the overall electronic properties of the interface. In addition, the charge transfer analysis reveals that an almost constant number of ∼0.35 electrons is transferred from BiOBr to MoS2, independent of the interface composition.
In agreement with optical analysis and DFT calculations, for the MoS2/BiOBr HJ we can derive the energy diagram illustrated in Fig. 4c. In pure 2D BiOBr, the photogenerated electrons recombine with the holes trapped by the defects and oxygen vacancies through photon emission. However, in the MoS2/BiOBr HJs, the VB depends on the layer ratio of the two contents according to our calculation. The HJ consisting of 1 layer of MoS2 and 1 layer of BiOBr has a fully localized VB over BiOBr, thus allowing the photogenerated holes to stay in BiOBr; while in the other two combinations, the VB is shared between MoS2 and BiOBr, which means the photogenerated holes are possibly dispersed both in MoS2 and BiOBr, and the number of transferred holes from BiOBr to MoS2 is dependent on the ratio of MoS2. DFT calculations prove that the holes photogenerated in BiOBr can remain in BiOBr or be partially transferred on the surface of MoS2. On the other hand, the lower conduction band minimum (CBM) of MoS2 allows the photoexcited electrons to transfer from BiOBr to MoS2. Overall, the system can be considered a Type-I HJ, in which a formal energy transfer (ET) process takes place from BiOBr to MoS2 following photo-excitation of the former through a charge exchange mechanism. Therefore, the charge recombination can proceed through two paths: the electron–hole recombination between the CBM and VBM of MoS2 or the recombination between the CBM of MoS2 and defects or oxygen vacancies on BiOBr (see the brown solid lines in Fig. 4c). As the amount of MoS2 is increased in the HJs, increasing amounts of excited electrons recombine with the holes inside MoS2, meanwhile the charges for the BiOBr-centred trap-assisted recombination are reduced.
The PEC HER performance of the prepared MoS2/BiOBr HJs was studied through linear sweep voltammetry (LSV) under neutral pH conditions (see the Experimental section for details). Acidic and basic conditions were not tested due to the pronounced instability of BiOBr in these media.43,44 The resulting polarization curves are reported in Fig. 5a. Before each LSV curve was recorded, 5 cycles of cyclic voltammetry with a scan rate of 50 mV s−1 over 0.2 to −0.6 V vs. RHE were performed in order to obtain a stabilized curve (Fig. S9†). From these curves it is evident that in the dark, the EC properties heavily depend on the composition of the HJs (dashed dotted curves). The addition of MoS2 nanosheets helps to decrease the onset potential systematically from ∼−524 mV to ∼0 mV going from the pure 2D BiOBr to 10% MoS2/BiOBr. The kinetics of the HER of different samples is compared by the Tafel slope, which is estimated from the polarization curves. As shown in Fig. 5b, the Tafel slopes of 2D BiOBr, 1% MoS2/BiOBr, 5% MoS2/BiOBr, 10% MoS2/BiOBr and 50% MoS2/BiOBr increase from ∼126 mV dec−1 to ∼245 mV dec−1, except for the 1% MoS2/BiOBr HJ, with a value higher than that found for 5% MoS2/BiOBr (see Table 1 for specific values). 2D MoS2, which is produced here in its thermodynamically stable phase, i.e. the semiconducting 2H-phase, does not activate the HER in the bias range of 0/−0.6 V vs. RHE, thus no Tafel slope is calculated. However, the PEC HER of these samples shows different trends from the EC HER (solid curves). The onset potential decreases from ∼−448 mV for 2D BiOBr to ∼281 mV for 1% MoS2/BiOBr, while it increases again when the weight ratio of MoS2 becomes 5%. From these data, it is apparent that the PEC HER performance is enhanced with respect to standard EC only for 2D BiOBr and the 1% MoS2/BiOBr HJ, while the performance in the PEC HER of 5% MoS2/BiOBr, 10% MoS2/BiOBr, 50% MoS2/BiOBr and 2D MoS2 is not better than the corresponding EC HER. The onset potential for the 10% MoS2/BiOBr HJ even increased from ∼0 V to ∼−336 mV, strongly indicating that light energy is not helpful in improving the HER activity in this sample. However, this unhelpful influence of UV light is probably due to the occurrence of a chemical transformation of the samples under UV irradiation, which can be deduced from the irreversible redox peaks in the initial CV curves (Fig. S9†).
Sample | Onset potential of EC HER (mV) | Onset potential of PEC HER (mV) | Tafel slope of EC HER (mV dec−1) | Tafel slope of PEC HER (mV dec−1) |
---|---|---|---|---|
2D BiOBr | −524 | −448 | 126 | 185 |
1% MoS2/BiOBr | −508 | −281 | 155 | 210 |
5% MoS2/BiOBr | −409 | −402 | 133 | 131 |
10% MoS2/BiOBr | 0 | −352 | 244 | 178 |
50% MoS2/BiOBr | −336 | −397 | 245 | 214 |
The influence of the light source on the kinetics of the reactions is also found to be different on samples with different compositions. In Table 1, a comparison of the values for each sample under EC and PEC conditions is presented, it is found that the Tafel slope is increased for both 2D BiOBr and 1% MoS2/BiOBr HJ, while decreased for 10% MoS2/BiOBr and 50% MoS2/BiOBr HJs. This indicates that the influence of UV illumination on the kinetics of the HER is different, depending on the composition of the samples. Moreover, the redox peaks that appear in these last samples in the range −0.2/0.1 V are possibly ascribed to the reduction peak of BiIII to Bi0.45 The photocurrent density–voltage curves of 2D BiOBr, 1% MoS2/BiOBr and 5% MoS2/BiOBr are shown in Fig. 5c. The photocurrent density of pure 2D BiOBr reaches 1.0 mA cm−2 at −0.6 V vs. RHE. 1% MoS2/BiOBr exhibits a noticeably enhanced photocurrent density of 1.9 mA cm−2 at −0.54 V vs. RHE, which then reduced to 1.5 mA cm−2 at −0.6 V vs. RHE. This is mainly because of a sluggish kinetic under the influence of the UV light source. Moreover, 1% MoS2/BiOBr shows an earlier start of the reaction under illumination compared to 2D BiOBr. However, the photocurrent is barely detected in the 5% MoS2/BiOBr or HJs with a higher ratio of MoS2.
Fig. 5d shows the transient photocurrent densities of the samples at the open circuit voltage. An enhancement of electron–hole separation for the 1% MoS2/BiOBr HJ is further confirmed by the increased photocurrent. Besides, a general photocurrent decay is observed in the three samples as well. The photocurrent of 2D BiOBr and 5% MoS2/BiOBr is negligible after two on/off light cycles, while 1% MoS2/BiOBr still shows a photocurrent density of around 1 μA cm−2 after three cycles. These results reflect better stability of the 1% MoS2/BiOBr HJ compared to the other samples.
To further elucidate the mechanisms of the PEC HER in the HJ samples, more characterizations were carried out on the aged electrodes to analyse their chemical and structural evolution. The samples after the PEC HER were collected for P-XRD measurements. As shown in Fig. 6a, the phase structure of 2D BiOBr is changed from BiOBr to mainly metallic Bi (ICSD number 64703) also with a broad peak at around 28.3° (the yellow dashed dotted line), which is a characteristic peak belonging to Bi2O3 (ICSD number 27151, see Fig. S10†). This result indicates that the BiOBr structure undergoes a chemical transformation: on one hand, the main part of Bi3+ is reduced to metallic Bi; on the other hand, Br is totally extracted during the PEC HER and part of the BiOBr is converted to Bi2O3. The BiOBr component in the HJs undergoes the same chemical transformation as 2D BiOBr except for the case of the 1% MoS2/BiOBr HJ, which mainly maintains the original BiOBr structure with a very small amount of reduced Bi. In addition, in order to probe the different evolution of the elemental composition and the surface chemical state in HJs, the aged 1% MoS2/BiOBr and 10% MoS2/BiOBr HJs were chosen for high-resolution XPS analysis. Fig. 6b displays the Br 3d spectra. The 1% MoS2/BiOBr HJ features the Br 3d peaks in the range from 67 eV to 70 eV, which is typical of Br in BiOBr.46 On the other hand, no signal for the 10% MoS2/BiOBr is detected, indicating the absence of Br on the surface of this electrode.47 Meanwhile, the lack of Br in the EDS spectrum proves that no Br is found below the surface as well (see Fig. S11†). The Bi 4f high-resolution XPS spectrum of 1% MoS2/BiOBr shows two peaks of Bi 4f5/2 (164.5 eV) and Bi 4f7/2 (159.1 eV), which derive from the lattice Bi(III) ions in BiOBr (Fig. 6c). However, the corresponding Bi 4f5/2 and Bi 4f7/2 peaks shifted to 164.2 eV and 158.8 eV for 10% MoS2/BiOBr. These shifts are likely due to the structural transformation from BiOBr to Bi2O3 in 10% MoS2/BiOBr.48 The existence of metallic Bi in 10% MoS2/BiOBr is further visible from the weak shoulder peak at 156.8 eV in the signal of Bi 4f7/2. Since XPS can only probe chemical information within less than 10 nm thickness of the sample, these results confirm that the surface of the 10% MoS2/BiOBr electrode can form a layer of Bi2O3, while the dominant metallic Bi is mainly formed under the surface. The O 1s spectra are shown in Fig. 6d. 1% MoS2/BiOBr after the PEC HER shows two peaks near 529.9 and 531.3 eV, attributed to O in BiOBr and O in chemisorbed species, and one small shoulder near 533.1 eV from the O vacancies on the surface.49,50 For the 10% MoS2/BiOBr after the PEC HER, the lattice O peak shifts to 529.6 eV. This is explained by the chemical transformation from BiOBr to Bi2O3, which is in accordance with the XRD and Bi 4f XPS spectra. Besides, the other two peaks can be ascribed to different types of chemisorbed O, like –OH and H2O.51 Therefore, we believe that after the PEC HER, the 1% MoS2/BiOBr HJ is the only sample able to maintain the original structure with only a small amount of BiOBr reduced to metallic Bi. Indeed it is able to sustain at least 16 cycles of CV under UV light (Fig. S9†). On the other hand in the 10% MoS2/BiOBr HJ (and likely also in the other ones), most of the BiOBr is reduced to Bi(0), leaving a Bi2O3 layer covering the surface of the photoelectrode.
Overall, the substantially different performance can be directly attributed to the different compositions between the 2D materials. As evidenced by the XRD and XPS analysis of samples after the PEC HER, different contents of MoS2 in the HJs result in different degrees of reduction of BiOBr during the PEC process. Specifically, pristine 2D BiOBr is able to absorb UV light and generate active species to boost the HER (see Fig. 5). However, this material is not stable under UV light because of the photoreduction from the photoelectrons. Our result is in accordance with the previous studies, which exhibited similar photocorrosion of 2D BiOBr. Surprisingly, building vdW HJ with a low ratio of MoS2 contributes to enhanced catalytic activity and structural stability, as proved in Fig. 5a, 6a and Fig. S12.† The reason may be that enhanced electron density at the MoS2/BiOBr interface is achieved due to the redistribution of the charge density, which is highly favourable for electrochemical reduction reactions such as HER. Meanwhile, the migration of extra photoelectrons to the conduction band of MoS2 and the accumulation of photogenerated holes on the VB of BiOBr can prevent BiOBr from being reduced (see Fig. 4c). However, higher contents of MoS2 in the HJs do not guarantee better efficiency and stability. With a higher amount ratio of MoS2 in the HJs, the assimilation of VBM of MoS2 and BiOBr allows the transfer of more photoexcited holes to MoS2, thus the extra electrons on BiOBr could facilitate the reduction of BiOBr (see Fig. S8c and S8d†). Besides, as the oxygen vacancies on the surface of BiOBr are able to trap photoexcited electrons, the consumption of these vacancies during the catalytic reaction might be another disadvantage to prevent BiOBr from photoreduction. As shown in Fig. S13† and Fig. 6d, oxygen vacancies are preserved during the PEC HER within 1% MoS2/BiOBr HJ but not within the 10% MoS2/BiOBr one.
To synthesize the HJs, multilayer MoS2 was first obtained by liquid phase exfoliation (LPE) with a tip-sonicator for 9 h using 500 mL of H2O and 5 g of MoS2 powder. The mixture was centrifuged at 2000 rpm for 20 min, and the suspension was collected and centrifuged at 8000 rpm for 5 min to precipitate the multilayer MoS2 as a precursor for the next step. Then, the as-synthesized BiOBr microspheres and multilayer MoS2 with different weight ratios were added to 50 mL NMP for further LPE with 9 h of tip-sonication to obtain heterojunctions with different weight ratios of MoS2 and BiOBr (see Table S1† for details). After centrifuging the mixture at 2000 rpm for 20 min, stable colloidal suspensions were obtained. Then, ethanol and H2O were added to precipitate the products.
The geometry optimization of the three interfaces was performed using the PBE-D3 functional with the VASP software.53–58 We used a plane wave basis set with an energy cutoff of 500 eV. The geometry optimization criteria were set to 10−5 eV for energy and to 0.02 eV Å−1 for forces on each atom. The calculations were performed at the Γ K-point, and a Gaussian smearing of 0.03 eV was applied for the whole calculations.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2nr04970h |
This journal is © The Royal Society of Chemistry 2023 |