Next Article in Journal
Chemoenzymatic Stereoselective Synthesis of trans-Flavan-4-ols via Lipase-Catalyzed Kinetic Resolutions
Previous Article in Journal
Insights into the Kinetics Degradation of Bisphenol A by Catalytic Wet Air Oxidation with Metals Supported onto Carbon Nanospheres
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thickness-Dependent Photocatalysis of Ultra-Thin MoS2 Film for Visible-Light-Driven CO2 Reduction

1
Institute of Atomic and Molecular Sciences, Academia Sinica, Taipei 10617, Taiwan
2
Center for Condensed Matter Sciences, National Taiwan University, Taipei 10617, Taiwan
3
Department of Chemistry, National Taiwan Normal University, Taipei 11677, Taiwan
4
Department of Materials Science and Engineering, National Taiwan University of Science and Technology, Taipei 10617, Taiwan
5
International Graduate Program of Molecular Science and Technology, National Taiwan University (NTU-MST), Taipei 10617, Taiwan
6
Molecular Science and Technology Program, Taiwan International Graduate Program (TIGP), Academia Sinica, Taipei 11529, Taiwan
7
Center of Atomic Initiative for New Materials, National Taiwan University, Taipei 10617, Taiwan
*
Authors to whom correspondence should be addressed.
These authors contributed equally.
Catalysts 2021, 11(11), 1295; https://doi.org/10.3390/catal11111295
Submission received: 30 September 2021 / Revised: 25 October 2021 / Accepted: 25 October 2021 / Published: 27 October 2021
(This article belongs to the Special Issue Photocatalysis in 2D Materials Science)

Abstract

:
The thickness of transition metal dichalcogenides (TMDs) plays a key role in enhancing their photocatalytic CO2 reduction activity. However, the optimum thickness of the layered TMDs that is required to achieve sufficient light absorption and excellent crystallinity has still not been definitively determined. In this work, ultra-thin molybdenum disulfide films (MoS2TF) with 25 nm thickness presented remarkable photocatalytic activity, and the product yield increased by about 2.3 times. The photocatalytic mechanism corresponding to the TMDs’ thickness was also proposed. This work demonstrates that the thickness optimization of TMDs provides a cogent direction for the design of high-performance photocatalysts.

1. Introduction

Photocatalytic CO2 reduction reaction (PC-CO2RR) is an elegant pathway in heterogeneous catalysis that transforms the CO2 molecule, which is a widely known pollutant and one of the major causes of global warming, into a variety of useful chemicals by solar light-driven conversion [1,2], and it is also expected to help achieve the goal of net-zero industrial emissions [3]. However, the materials that have so far been investigated under the PC-CO2RR principle still show poor conversion efficiencies and are still far from satisfying practical applications [4]. One of the main issues is the lack of high efficiency in photocatalysts that can be engaged in CO2 reduction. Although a large number of metal oxide-based catalysts that can respond to ultraviolet light (such as titanium dioxide (TiO2)) have been reported to be suitable for PC-CO2RR applications, most of these materials still possess relatively low conversion efficiencies [5,6]. Consequently, researchers have devoted a great deal of time and effort to developing visible-light-active photocatalysts with appreciable conversion efficiencies [7,8]. In addition to the issue of the absorption of a broad range of wavelengths present in sunlight, another major factor that enhances the performance of a photocatalyst is the ability of excellent carrier separation and transportation. Presently, the conundrum faced by researchers in this field is that the low charge separation efficiency of the photocatalyst leads to the rapid recombination of photogenerated electrons (e) and holes (h+) in the entire photocatalytic process, making the conversion efficiency very poor [9]. Therefore, a myriad of strategies have been explored to manipulate the catalyst’s charge transfer and/or spatial separation capabilities to obtain higher catalytic activity [7,8,9,10], including doping [11], co-catalysts, reduced particle size, and heterojunctions.
Two-dimensional transition metal dichalcogenide (TMD)-based catalysts have been predicted to be promising candidates for improving photocatalytic hydrogen evolution and CO2 conversion efficiency in recent times [7,12,13], with several glowing mentions of layered molybdenum disulfide (MoS2). Many reports indicate that MoS2 shows great promise as a catalyst due to its tunable optical characteristics, suitable band-gap potential, and high stability under continuous light illumination, allowing it to perform visible-light photocatalysis [7,12,13,14]. For instance, MoS2 exhibits a good ability to convert CO2 into valuable fuels (CH4, CO, CH3OH, HCOOH) through PC-CO2RR [15]. As a visible light-driven photocatalyst for PC-CO2RR processes, it is necessary to conduct a thorough and systematic study of the changes in light absorption and charge recombination caused by tuning the number of layers of MoS2 and the subsequent photocatalytic activity. It has been widely reported that a controlled number layers of MoS2 nanosheets can be made by mechanical and chemical exfoliation methods [16,17]. When the size of MoS2 nanosheets is reduced, it will not only change the ratio of the length of the edge to the base surface of the MoS2 nanosheets but will also change the surface chemistry [16,17] of the material. In the past, many publications have reported that the edge sites in MoS2 are more active than the basal plane (which is relatively inert) [17,18], with respect to catalytic activity. Therefore, it is understandably challenging to choose a method of MoS2 production that can avoid excessive edge exposure while only tuning the number of layers in the catalyst to solve the above-mentioned problems. The surface properties and edge size of the thin film-type photocatalyst can remain unchanged when the film thickness changes. In this way, the influence of thickness on the overall catalytic performance can be isolated from other convoluting factors.
In this study, we have successfully synthesized ultra-thin MoS2 films (MoS2TFs) with different thicknesses as model catalysts and investigated their PC-CO2RR activities under visible light. We demonstrated that the layer numbers and crystallinity of MoS2TFs can be fine-tuned by a three-step post-sulfurization and chemical vapor deposition process. The results indicate that the photocatalytic activity was significantly enhanced with the increasing thickness of the MoS2TFs. Impressively, 25 nm MoS2TFs exhibit the highest photocatalytic activity because they have excellent optical absorption and an appropriate grain size. This strategy not only provides a new perspective for enhancing photocatalytic activity by controlling the thickness of the film but also gives us a deeper understanding of the mechanism of optical absorption and charge separation in promising layered photocatalysts such as MoS2.

2. Results and Discussion

2.1. Synthesis of Photocatalytic Ultra-Thin Film

In this paper, we report the growth and characterization of MoS2TFs on sapphire substrates, wherein we used thermal evaporation in combination with chemical vapor to synthesize the thin films to be studied. This method is similar to the one that was reported in our previous publication [19]. Herein, we provide a detailed report on the actual preparation of the MoS2TFs. We also elaborate upon the methods that can be utilized to control and identify the disorder that could be observed in the thin film that we synthesized. Figure 1a depicts the schematic diagram of the three-step method proposed in this study. The first step is the deposition of the precursor MoO3 film on the c-sapphire substrate by thermal evaporation. MoO3 films prepared by thermal evaporation have the distinct advantages of adjustable thickness, large area, continuous film growth, good uniformity, and high quality. The second step involves annealing the MoO3 film using a gaseous mixture of Ar and H2 at 500 °C for a total duration of 1 h. In this step, the MoO3 film underwent partial reduction and was converted to an MoO2 film, which is a form that is easier to convert to our target MoS2TFs. The final step is the sulfurization of the MoO2 films using a gaseous mixture of Ar and H2S at 900 °C for 0.5 h to produce the target product (MoS2TFs). Figure 1b shows the temperature profile of the MoS2TFs growth process for hydrogenation and sulfurization steps. Figure 1c shows the optical photographs of the c-sapphire substrate before and after the growth of MoS2 thin films. From this picture, it can be clearly seen that the MoS2TFs fully cover the entire sapphire substrate and have excellent uniformity. In addition, the color of the sample surface changes from transparent to dark yellow, as there is an increase in film thickness.

2.2. Characterization of Photocatalytic Ultra-Thin Film

Figure 2a–d illustrate the thicknesses and surface roughness of the MoS2TFs measured by atomic force microscopy. All specimens for the thickness measurements were prepared by scratching the TFs. The area that was scanned in the AFM measurements is 10 × 10 µm for these MoS2TF samples (see the first column in Figure 2). The height profiles of the AFM images (second column in Figure 2) show that the thickness of the four MoS2TFs samples was 7, 17, 25, and 50 nm, respectively. In addition, the surface roughness (Rq value) of the four MoS2TFs samples was between 0.5 and 2.7 nm (third column in Figure 2). From these AFM results, it can be conclusively stated that the MoS2TFs prepared by this three-step method are layer-controlled, have excellent homogeneity, and have very smooth surfaces.
Confocal Raman spectroscopy was used to further characterize the quality and uniformity of synthesized MoS2TFs with various thicknesses. Figure 3 shows the Raman spectra of the 25 nm MoS2TFs on sapphire substrates synthesized under different reaction temperatures of 500 °C, 700 °C, and 900 °C, respectively. Figure 3a shows the two main characteristic Raman peaks of MoS2 for all of the MoS2TFs samples, namely the in-plane E 2 g 1 peak at ≈384 cm−1 and the out-of-plane A1g peak at ≈409 cm−1. In addition, a small LA(M) sub-peak at ≈227 cm−1 related to the defects in MoS2 can also be observed, which is attributed to the longitudinal phonons at the M point in the Brillouin zone [20,21,22]. The peak at 418 cm−1 comes from the sapphire substrate (A1g mode) [23]. Furthermore, the intensity ratio of the LA(M) sub-peak to the A1g peak can be considered to be a marker for evaluating the quality of MoS2TFs. In general, the good quality MoS2 samples with lower disorder can be determined by lower LA(M)/A1g values [24,25]. In order to elucidate our point, we plotted the LA(M) to A1g peak intensity ratio of the MoS2TFs samples as a function of the reaction temperature, as shown in Figure 3b. The result shows that when the reaction temperature increases (from T = 500 to 700 °C), the LA(M)/A1g intensity ratio shows a downward trend (from 0.15 down to 0.1). No clear LA(M) peak was identified as the reaction temperature approaches 900 °C (Figure 3a). To the best of our knowledge, in comparison to previous studies [24,25], the LA(M) to A1g peak intensity ratios reported in this study are the lowest values to date. These data clearly indicate that the MoS2TFs samples grown by the high-temperature process described herein have better film quality.
Figure 4a presents the typical Raman spectra of the prepared MoS2TFs with thicknesses increasing from 7 to 50 nm. The Raman spectra of commercial monolayer MoS2 on sapphire substrate and bulk MoS2 crystal is also shown in Figure 4a, for ease of comparison. Raman spectra of the monolayer MoS2 clearly showed the two characteristic peaks at 406.5 cm−1 (A1g mode) and 386.1 cm−1 (E12g mode). The separation between these two peaks (A1g and E12g modes) is 20.4 cm−1. According to previous reports [20,21], the frequency difference between the two main modes (Δk) could be used to identify the number of MoS2 layers. When the number of MoS2 layers increases, the E12g mode will display a red-shift, whereas the A1g mode undergoes a blue-shift [20,21]. It is worth noting that the incremental shift becomes smaller and smaller as the number of layers increases. When the number of MoS2 layers increases, the E12g mode will display a red-shift, whereas the A1g mode undergoes a blue-shift [20,21]. It is worth noting that the incremental shift becomes smaller and smaller as the number of layers increases. When the number of MoS2 layers is four or more, the frequencies of these two modes will approach the values for bulk MoS2 [26]. In Figure 4a, we can clearly observe that when the MoS2TF thickness increases from 7 nm up to 50 nm, there is no main peak shift phenomenon that is apparent in the Raman spectra of these MoS2 samples, i.e., the two main Raman peaks (A1g and E12g modes) of these MoS2TF samples were observed at ~410 cm−1 and ~384 cm−1, respectively. In addition, we also found that the two main Raman peak positions for all MoS2TF samples are close to the Raman peaks of commercial bulk MoS2 crystals, and there is a slight shift (MoS2TFs has a blue shift (~1 cm−1) compared to the bulk MoS2), as shown in Figure 4b. It is worth noting that the few-Layer MoS2 films (4 to 6 layers) synthesized by mechanical exfoliation didn’t present this slight shift phenomenon [26]. We assume it may be related to the fact that MoS2TF is grown on sapphire substrates, instead of being directly peeled off from the bulk MoS2 crystal. In addition, it has been reported that due to factors inherent to the substrate (such as, deformations and mismatch in the coefficient of thermal expansion), the single-layer and double-layer TMD materials will generate strain, which causes these two Raman modes to shift slightly [27,28]. The detailed mechanism of the onset of shifts in the Raman peaks in multilayer MoS2 or the ultra-thin films synthesized by us on sapphire substrates is still unclear, and further investigation is needed.
In order to evaluate the uniformity of the MoS2TFs prepared by our process, we measured the Δk value at 16 different points on the MoS2TFs with increasing thickness, as shown in Figure 4c. The results show that the measured Δk values of all MoS2TF samples are between 25.4 to 25.9, which are close to the values of bulk MoS2 crystals. This result demonstrates the excellent uniformity of our custom-grown MoS2TFs with various thicknesses. All Raman results also indicate that uniform thickness-controlled MoS2TFs can have the advantage of large scalability of growth on sapphire substrate.
Evolution of the optical properties of the as-prepared MoS2TFs was further investigated by measuring the optical absorbance spectra of MoS2TFs as a function of thickness. Figure 5a shows the UV-Vis-NIR absorption spectra of 7, 17, 25, and 50 nm MoS2TFs, respectively. All deposited films exhibit four prominent peaks characteristically indicating 2H-MoS2 excitonic features in the visible wavelength region. More specifically, the A and B excitons that form at the K-point of the Brillouin zone appeared in the 600 to 700 nm wavelength region, while the C and D excitonic peaks were located in the wavelength range from 400 to 450 nm. The absorption spectrum of the sapphire substrate has been shown in Figure 5a. Looking at Figure 5b we found that with increasing thickness of MoS2TFs from 7 to 25 nm, the absorbance at peak of A exciton increased rapidly from 0.25 to 0.6. However, when the film thickness increased to 50 nm, the MoS2TFs showed a saturation trend of absorbance, similar to the absorbance spectrum of 25 nm film. The absorbance at peak of B exciton presents a trend similar to the A exciton. We also observed that the A exciton energy (EA) of MoS2TFs on sapphire samples will decrease as film thickness increase (from 1.851 eV to 1.834 eV), and it will converge when the thickness reaches 25 nm (about 40 layers), as shown in Figure 5c. It is worth noting that these results are not the same as previously reported in literature [29]. In a previous study, the A exciton position of the micromechanically-exfoliated nanosheets was seen to decrease from 1.89 to 1.83 eV as the thickness increases from 1 to 6 layers. When the number of layers reaches 5 or 6, the downward trend of A exciton position begins to converge. In addition, in the case of liquid-phase exfoliated nanosheet samples, when the number of layers reaches 10 layers, the exciton peak position will begin to converge. We think this may be related to the difference between the dielectric constant of the CVD-grown MoS2TFs samples and the exfoliated samples. Since the interlayer electronic hybridization and dielectric shielding effect was changed as a function of the number of MoS2 layers, the resulting properties of A excitons also changed comparably. However, the detailed mechanism of the position shift pertaining to the energy of the A exciton of MoS2TFs on sapphire substrates is still unclear. Therefore, we suggest that the dielectric constant of MoS2TFs grown on a sapphire substrate should be measured directly using spectroscopic ellipsometry combined with quantum electrostatic heterostructure (QEH) model calculation [29]. This approach can help to elucidate the layer-related exciton effect of MoS2TFs more quantitatively.
The optical band gap of a photocatalyst can be calculated by the following formula: (αhv)1/n = A(hν-Eg) [30,31]. In this equation, ‘α’ is the absorption coefficient, ‘h’ is Planck’s constant, ‘υ’ is the photon frequency, ‘A’ is a constant, and ‘Eg’ is the band-gap energy. The ‘n’ is the exponent term, which denotes the nature of the electronic transition. When n = 1/2, it signifies a direct allowed transition, and when n = 2, it represents an indirect allowed transition. According to previous reports [32], monolayer MoS2 is a direct energy band-gap semiconductor, and an increase in the number of layers (including multi-layer, film, and bulk crystal) causes a shift of behavior to an indirect band-gap semiconductor. Not only that, the Eg value of MoS2 will change from 1.9 eV for a monolayer to 1.2 eV for bulk crystal. Tauc plots as a function of photon energy for MoS2TFs can be used to determine the Eg value by extrapolating the straight-line portion of the indirect electronic transitions, as shown in Figure 6. Here, we used the indirect transition value (n = 2) to calculate the Eg value of all MoS2TF samples (Figure 6a–d). The results show that when the film thickness is increased from 7 to 50 nm, the Eg value of MoS2TFs will decrease from 1.71 to 1.60 eV (Figure 6e). The Eg values of monolayer and bulk crystal also shown in Figure 6e for reference purposes.
Photoluminescence (PL) is another optical spectroscopic technique used to measure the band gap of semiconductors. Monolayer MoS2 is a direct bandgap semiconductor that exhibits significant PL intensity and prominent excitonic features (A and B exciton peaks) in PL spectra [33]. Multi-layer MoS2 is an indirect band gap semiconductor, and its PL intensity will decrease rapidly as the number of layers increases. In bulk crystal MoS2, the PL intensity is relatively weak; so much so, that in some cases no PL excitation at all is observable. Figure 7 shows the PL spectra of MoS2TFs on sapphire with various thicknesses. The results show that our as-deposited MoS2TFs do not exhibit strong PL intensity. When the thickness of MoS2TFs is increased from 7 to 50 nm, the PL intensity will decrease dramatically (Figure 7a). No clear A exciton peak was identified in the PL spectra of 25 nm and 50 nm MoS2TFs. The prominent MoS2 Raman modes were observed at around 550 nm wavelength region for all MoS2TFs samples. Not only that, the broad PL spectra profile and non-obvious excitonic features (A and B exciton peak) of the MoS2TFs are similar to bulk crystal MoS2 (Figure 7b). The PL signal of the sapphire substrate is still visible and overlap with A exciton peak of MoS2 occurs at a wavelength of around 690 nm (Figure 7a,c). Therefore, it is difficult to determine the position of the A exciton peak and bandgap for all MoS2TFs samples. The PL spectra of our as-prepared MoS2TFs exhibited similar excitation phenomena as chemically exfoliated MoS2 layers. In the case of chemically exfoliated MoS2 layers [33], when the thickness of the MoS2 layer is greater than 7 nm, no clear A exciton peak can be observed in the PL spectrum.
X-ray diffraction (XRD) was used to study the structural properties and crystalline parameters of the as-prepared MoS2TFs. From the XRD patterns shown in Figure 8, it can be observed that all MoS2TF samples are unequivocally identified as 2H MoS2 with a strong characteristic peak appearing at 2θ = 14.3°, which was assigned to the MoS2 (002) lattice plane [34]. However, the peak denoting the (004) lattice plane of MoS2 at 2θ = 29° cannot be observed in our MoS2TF samples. Furthermore, we used the Scherrer Equation to analyse the grain size of the MoS2TFs, as listed in Table 1. According to the Scherrer formula, D = Kλ/(β cos θ) [35], where ‘D’ is the average grain size in vertical direction of crystal structure, ‘K’ is the shape factor having a typical value of about 0.9, ‘λ’ is the wavelength of the X-ray source, ‘β’ is the full width at half maximum intensity of the peak (in Rad), and ‘θ’ is the Bragg angle. As seen in Figure 8b, we found that as the thickness of MoS2TFs increased from 7 to 25 nm, the grain size of the MoS2TFs increased from 5.6 to 21 nm. However, when the film thickness increased to 50 nm, the grain size of MoS2TFs was constrained to around 25 nm. This data indicates that all the MoS2TF samples have good crystallinity, and the grain size is almost identical to the measured film thickness, except in the case of the 50 nm sample.

2.3. Photocatalytic Performance

The evaluation of the PC-CO2RR performance of the MoS2TFs with various thicknesses was performed in a visible light-driven gas-phase photoreaction vessel for 24 h, as shown in Figure 9. Briefly, the MoS2TFs with an area of 2 cm2 were placed at the center of the photoreactor. Then, N2 gas and moist CO2 gas were sequentially passed through the photoreactor for 30 min and 1 h, respectively. A 100 W Xe solar simulator with an AM1.5 filter was used as the light source to study the photoconversion of CO2 and water vapor over a time period of 24 h. The schematic representation of our gas-phase and batch-type PC-CO2RR system is shown in Figure 9a. In order to avoid the overestimation of the catalytic activity of our samples due to carbon contamination on the surface of the reactor, we carried out two blank tests as the background for all MoS2TF samples (Supporting Figure S1 in Supplementary Material). This background value will be subtracted from the spectral data of the final gas products produced by the MoS2TF catalysts after 24 h of photocatalysis. Figure 9b,c show the product yields of the MoS2TFs with different reaction temperatures and different thicknesses. All the MoS2TFs samples are capable of catalyzing the reduction of CO2 into methane (CH4), ethylene (C2H4), acetaldehyde (CH3CHO), and acetone (C3H6O). Methane and acetaldehyde were identified as the first and second major products, respectively. In addition, small amounts of ethane and acetone were also detected. The corresponding gas production yield and quantum efficiencies (QE) of all MoS2TF samples are summarized in Table 2 and Table 3, respectively.
The QE of the MoS2TFs was determined as the ratio of the effective electrons used for gas production, such as the CH4 molecule, to the total input photon flux, and it was calculated by the following equation [36,37]:
QE % = Effective   electrons Total   photons × 100 % = 8 × Y × N Θ × T × S × 100 %
where ‘Y’ is the product yield of CH4 (mol), ‘N’ is Avogadro’s number (6.022 × 1023 mol−1), ‘Θ’ is the photon flux (2.46 × 1017 cm−2 s−1), ‘T’ is the reaction time (s), and ‘S’ is the area of illumination (cm2). The area under illumination (S) of the MoS2TF samples is 2 cm2. For example, the QE% of CH4 for 25 nm MoS2TFs after 24 h of PC-CO2RR can be calculated as QE = (8 × 1.38 × 10−9 × 6.022 × 1023)/(2.46 × 1017 × 24 × 3600 × 2) × 100% = 0.000016%. Based on the same calculation method, the QE%s of C2H4, CH3CHO, and C3H6O for the 25 nm MoS2TFs were calculated to be 0.000015%, 0.000033%, and 0.0000047%, respectively. The total QE of all gas products for MoS2TFs is 0.000068%, and it is calculated by the following equation:
Total QE% = QE% of CH4 + QE% of C2H4 + QE% of CH3CHO + QE% of C3H6O.
Figure 9b shows that the 25 nm MoS2TFs prepared at a reaction temperature of 900 °C can convert the highest amount of CO2 into the four gas products (total amount of gas products is 4.79 nmol/cm2), which is followed by the samples produced at a reaction temperature of 700 °C (3.16 nmol/cm2), and these in turn are followed by the MoS2TF samples synthesized at a reaction temperature of 500 °C (2.09 nmol/cm2). The resulting total QEs for the MoS2TF samples produced at reaction temperatures of 500, 700, and 900 °C were 0.000029%, 0.000045%, and 0.000068%, respectively. In Figure 9c, we observe that as the thickness of MoS2TFs increased from 7 to 25 nm, the gas production yield and total QE exhibited a gradually increasing trend (2.13 to 4.79 nmol/cm2 and 0.000030 to 0.000068%). However, when the film thickness increased to 50 nm, the gas production yield and QE showed a downward trend (4.16 nmol/cm2 and 0.000059%). These results indicate that the 25 nm MoS2TFs exhibited the highest photocatalytic activity.

2.4. Photocatalytic Mechanism

From the above-mentioned CO2 reduction activities of different MoS2TF photocatalysts, we can conclude that the crystallinity and thickness of MoS2TFs play an important role in the PC-CO2RR processes. We will try to understand the underlying photocatalytic mechanism from the following aspects, as illustrated in Figure 10.

2.4.1. Size Effect

In case 1 as illustrated in Figure 10, the typical process of PC-CO2RR on a semiconductor-based photocatalyst has been schematically depicted. Briefly, the PC-CO2RR process consists of five consecutive steps: light absorption, charge separation, CO2 adsorption, surface redox reaction, and product desorption [38,39]. These five steps can be further divided into two theoretical categories, that is, photophysics (the first two steps) and photochemistry (the next three steps). For the purpose of understanding the performance of the catalysts we have synthesized, we will mainly focus on understanding the relationship between photophysics and the activity of our photocatalysts. The first step shows that it is necessary to analyze the ability of photocatalysts to absorb photons and consequently generate electron–hole pairs, which is acutely affected by factors such as the energy band gap (Eg) and thickness of the photocatalyst. The second step hinges on the ability of photocatalysts to effectively separate the photogenerated electrons and holes, which is directly related to charge recombination phenomena. The crystallinity and surface properties of photocatalysts all tangibly affect charge recombination. Therefore, in order to improve the overall photocatalytic efficiency, it is necessary to increase the optical absorption of the incoming visible light radiation, to generate electron–hole pairs with vigor, efficiently separate the photo-generated charge carriers, and suppress their recombination, as shown in Figure 10.
It is worth noting that the utilization of the so-called “nanosizing strategy” has proven to be very beneficial toward improving the performance of visible light-driven photocatalysts [10,40]. For instance, Ta3N5 nanoparticles with a particle size of 30–50 nm exhibited higher photocatalytic activity for H2 evolution, as compared to traditional bulk Ta3N5 particles (300–500 nm) [40]. A photocatalyst with a smaller particle size has the advantage of a higher surface area, which also increases the density of surface catalytic sites. However, it usually results in lower crystallinity and may reduce catalytic activity. Moreover, in particle type photocatalysts with a three-dimensional geometry, calculative complications are caused by particle size reduction, such as crystal facet and morphology change. This makes it a challenging task to lucidly explain the actual catalytic mechanism. Therefore, we propose the utilization of thin-film type photocatalysts with a two-dimensional geometry to isolate the “thickness effect” and avoid the above problems for the study of specific catalytic mechanisms study.

2.4.2. Effect of Optical Absorption

According to past reports [32,41], by controlling the number of layers of MoS2, not only can the bandgap be fine-tuned, but the light absorption performance can also be optimized. In a previous study, Atwater’s group reported that the sub-15 nm thick TMD flakes created by mechanical exfoliation, have near-unity optical absorption [41]. Their calculated and measured absorption spectra showed that the ultra-thin (14–17 nm) MoS2 flakes on an Ag back reflector have the highest broadband absorption. According to our experimental data, the energy bandgap of MoS2TF samples with different thicknesses is about 1.7 to 1.6 eV and there is not too much variation in these values (Figure 6e). However, there are obvious differences in the optical absorption of MoS2TF samples with different thicknesses (Figure 5b). The ultra-thin MoS2TFs on sapphire with a thickness of 25 nm exhibited the highest optical absorption. This result is similar to the optical absorption capacity of MoS2 flakes reported in literature. Therefore, we believe that the sufficient optical absorption of visible light radiation by films having an optimized thickness of 25 nm (sample (case II)), which exhibits better photocatalytic activity than 7 or17 nm MoS2TFs samples (case III), as shown in Figure 10.

2.4.3. Effect of Grain Size

According to our experimental data, the grain size of MoS2TFs samples with different thicknesses is about 7 to 25 nm. In addition, a point to be noted is that the 25 nm MoS2TF sample has almost the same thickness as the grain size. The 50 nm MoS2TF sample has grain boundaries because the grain size is only 25 nm. Therefore, we propose a possible reason to explain why the 25 nm MoS2TFs sample (case II) has better photocatalytic activity than 50 nm MoS2TFs samples (case IV), as shown in Figure 10. It is well known that the grain boundary of semiconductors is the primary site of charge recombination, and changes therein significantly affect the charge transport [42,43]. Based on this knowledge, we think it is reasonable that the grain size of ultra-thin MoS2TFs samples is similar to film thickness (case II: D is similar to t), which can promote the internal photogenerated carriers to migrate to the surface and efficiently drive the catalytic reaction (case II). On the contrary, the grain size of thick MoS2TFs samples is smaller than the film thickness (case IV: D is smaller than t), resulting in the formation of grain boundaries inside the film and the subsequent suppression of the probability of photogenerated carrier transfer to the surface. We studied two additional thick MoS2TFs samples and measured their grain size and catalytic activity to support this scientific point (Supporting Figures S2 and S3 and Supporting Table S1). The results clearly indicate that when the thickness is continuously increased (from 40 to 60 nm) without increasing the grain size (maintained at around 20 nm), the photocatalytic activity does not increase dramatically. In our current TFs process, the biggest grain size of MoS2TFs with a thickness of 7 to 60 nm is about 20 nm, and the 25 nm MoS2TFs sample has the best photocatalytic activity of about 4.79 nmol/cm2. Therefore, we believe that in addition to light absorption, the grain boundary inside the photocatalyst is another important factor that significantly affects the catalytic activity.

2.4.4. Stability Test

In order to test the stability of our MoS2TFs, this model catalyst was reused in a fresh reactor for three successive cycles. According to Supporting Figure S4, the total gas production yield slightly decreases over time, and the reduced ratio was calculated to be about 10% after three cycles. The slight decrease in the activity after the three cycles might be due to the inactivation or poisoning of the catalytic sites on the basal plane MoS2TFs. In addition, we also found a slight change in the production ratio of methane and acetaldehyde (methane production is increased and acetaldehyde formation is decreased). The results indicated that when reducing the catalytic sites on the MoS2TFs surface, the catalytic activity and the reaction pathway of the product will be consequently altered. The detailed reaction mechanism is still unclear, and further investigation is needed.
In addition to optical absorption and the grain size of the MoS2TFs, other related phenomena worth studying further include measuring the photocarrier recombination losses, surface reaction investigations, product analysis, and isotopic labeling. For recombination, it is worth noting that traditionally, time-resolved absorption and PL spectroscopies can successfully study the dynamic behavior of photocarriers in photocatalysis [44,45]. However, when the thickness increases, neither our as-prepared MoS2 film nor the MoS2 flake [29] have any obvious PL emission. Therefore, this tool cannot be used to understand the behavior of carrier recombination in the case of our MoS2TFs. Other alternative tools need to be selected and developed, such as advanced space-resolved techniques [46]. For surface reactions, the relationship between the surface of TMD catalysts and the surface reaction pathways has still not been clearly understood. It is necessary to use in situ catalytic tools to study the surface reactions on the catalyst surface, such as ambient pressure X-ray photoelectron spectroscopy for CO2 adsorption [47] and infrared absorption spectroscopy for surface species [48]. For product analysis, the accurate measurement of gas products relies heavily on the sensitivity and number of gas analysis detectors (such as gas chromatography-thermal conductivity detector (GC-TCD) and chromatography-flame ionization detector (GC-FID)) [49]. In most cases, only one gas detector is used, which often leads to inaccuracies in the overall gaseous product evaluation. For instance, H2 could also be one of the products of CO2 reduction, and its detection requires GC-TCD. Not only that, the H2 production rate should also be factored into the quantum efficiency calculations. The complete analysis of the products of CO2 reduction with the appropriate equipment is one of the directions that certainly must not be overlooked. While isotopic labeling technology may be a slightly more expensive alternative to confirm the accuracy of the products and product-specific active sites [50], it is certainly a project well worth being further investigated.

3. Materials and Methods

3.1. Sample Preparation

Uniform MoS2TFs on (0001) sapphire substrate with different thicknesses were prepared by two-step post-sulfurization of vacuum-deposited molybdenum trioxide (MoO3) films. Initially, a c-face sapphire substrate was cleaned with ethanol, acetone, and DI water for 5 min, respectively. Then, the clear substrates were subjected to O2 plasma treatment for 10 min prior to the deposition of the MoO3 film. The precursor MoO3 powder was supplied by Alfa Aesar with 99.95% purity, which was deposited on the top of sapphire substrate as a thin film, using thermal evaporation. In this process, the desired thickness of the film to be deposited can be closely monitored using a quartz crystal microbalance (QCM) by measuring the change in frequency of a quartz crystal resonator. The deposition rate was maintained in the range of 0.1 to 0.2 kÅs−1. Thereafter, all samples were placed into a chemical vapor deposition (CVD) system to perform the post-sulfurization process. Before introducing H2S gas into the system, the MoO3 films were annealed at 500 °C for 1 h under vacuum in an Ar-H2 environment (4:1), with the aim of reducing them into MoO2. Finally, the CVD growth process was performed at atmospheric pressure with 100 sccm of Ar and an H2-H2S mixture (flow rate 1:4). The post-sulfurization process was performed at 1000 °C at the rate of 30 °C min−1, and this condition was maintained for 30 mins to ensure successful sulfurization of the MoS2TFs.

3.2. Characterization

The as-grown MoS2TFs on sapphire substrates were systematically characterized by the following microscopy and spectroscopy-based tools. Firstly, the thickness and surface roughness of MoS2TFs was examined using the tapping mode of atomic force microscopy (AFM) with a scanning rate of 0.5 Hz and 10 µm scanning area using a Bruker Dimension Icon Atomic Force Microscope. The Raman and photoluminescence (PL) spectra were recorded on a Jobin-Yvon LabRAM H800 system with a 532 nm Nd:YAG laser as the excitation source and a spot size of ≈1 μm. In addition, the optical absorbance of the MoS2TFs was measured by a UV-Vis-NIR spectrophotometer (JASCO V-670). The optical band gap of the MoS2TFs was determined using Tauc plots. After that, the crystal structure and grain size of the MoS2TFs was determined using a Bruker D2 PHASER X-ray diffractometer.

3.3. Photocatalytic Activity Measurement

The gaseous product species generated by all the MoS2TF photocatalysts after the PC-CO2RR process were analyzed by gas chromatography (GC), on an Agilent 6890 system using a glass PLOT column (RT-Q-BOND), and a flame ionization detector (FID).

4. Conclusions

In summary, the lab-grown MoS2TFs on sapphire with thickness from 7 to 50 nm were used as a model catalyst and the photocatalytic CO2 reduction activities of thickness dependent MoS2TFs were systematically investigated. The MoS2TFs with 25 nm thickness exhibited the highest photocatalytic activities under visible light irradiation, corresponding to a gas production yield of 4.79 nmol/cm2 with a QE of 0.000068%. It is demonstrated that 25 nm MoS2TFs have the best ratio of light absorption and grain size, as compared to films of other thicknesses. We also proposed a systematic photocatalytic mechanism to relate the dependence of the thickness of TMD based photocatalysts, on their eventual catalytic mechanism. This study provides a novel outlook towards the development of high-efficiency two-dimensional material catalysts for the efficient reduction of CO2 into other useful chemical species.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11111295/s1, Figure S1: The total gas products of two kinds of blank test conditions for photocatalysis: 1.47 nmol/cm2 for blank test-1 and 0.27 nmol/cm2 for blank test-2, respectively, Figure S2: XRD pattern of the MoS2TFs on sapphire with different thicknesses: (a) t = 40 nm and (b) t = 60 nm, Table S1: The calculated grain size of MoS2TFs on sapphire from (002) peak using Scherrer’s formula, Figure S3: The PC-CO2RR activity of MoS2TFs on sapphire with thickness increasing from 40 to 60 nm, Figure S4: The stability study of the photocatalytic CO2 reduction over the MoS2TFs with 25 nm thickness.

Author Contributions

Y.-F.H. conceived and designed the experiments. K.-W.L. and F.R.Z.F. carried out the thin-film deposition and material characterizations. K.-W.L., F.R.Z.F., Y.-F.H., S.-H.T., S.-W.K. and C.-H.W. performed the GC experimental setup, calibration, photocatalytic activity, and data analysis. Y.-F.H., V.A.M. and K.-H.C. co-wrote the manuscript. K.-H.C. and L.-C.C. supervised the project. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Ministry of Science and Technology, Taiwan (Grant Numbers: 108-2119-M-002-030, 109-2123-M-002-004, 110-2123-M-002-006, and 108-2112-M-001-046-MY2) and by Academia Sinica (Grant Number AS-iMATE-110-34). Y.-F.H. acknowledges support from IAMS Junior Fellowship by Institute of Atomic and Molecular Sciences, Academia Sinica, Taiwan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kätelhön, A.; Meys, R.; Deutz, S.; Suh, S.; Bardow, A. Climate change mitigation potential of carbon capture and utilization in the chemical industry. Proc. Natl. Acad. Sci. USA 2019, 116, 11187–11194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Hepburn, C.; Adlen, E.; Beddington, J.; Carter, E.A.; Fuss, S.; Mac Dowell, N.; Minx, J.C.; Smith, P.; Williams, C.K. The technological and economic prospects for CO2 utilization and removal. Nature 2019, 575, 87–97. [Google Scholar] [CrossRef] [Green Version]
  3. Davis, S.J.; Lewis, N.S.; Shaner, M.; Aggarwal, S.; Arent, D.; Azevedo, I.L.; Benson, S.M.; Bradley, T.; Brouwer, J.; Chiang, Y.M.; et al. Net-zero emissions energy systems. Science 2018, 360, 6396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Bushuyev, O.S.; De Luna, P.; Dinh, C.T.; Tao, L.; Saur, G.; van de Lagemaat, J.; Kelley, S.O.; Sargent, E.H. What should we make with CO2 and how can we make it? Joule 2018, 2, 825–832. [Google Scholar] [CrossRef] [Green Version]
  5. Indrakanti, V.P.; Kubicki, J.D.; Schobert, H.H. Photoinduced activation of CO2 on Ti-based heterogeneous catalysts: Current state, chemical physics-based insights and outlook. Energy Environ. Sci. 2009, 2, 745–758. [Google Scholar] [CrossRef]
  6. Chang, X.; Wang, T.; Gong, J. CO2 photo-reduction: Insights into CO2 activation and reaction on surfaces of photocatalysts. Energy Environ. Sci. 2016, 9, 2177–2196. [Google Scholar] [CrossRef]
  7. Voiry, D.; Shin, H.S.; Loh, K.P.; Chhowalla, M. Low-dimensional catalysts for hydrogen evolution and CO2 reduction. Nat. Rev. Chem. 2018, 2, 0105. [Google Scholar] [CrossRef]
  8. Djurišić, A.B.; He, Y.; Ng, A.M. Visible-light photocatalysts: Prospects and challenges. APL Mater. 2020, 8, 030903–030924. [Google Scholar] [CrossRef] [Green Version]
  9. Marszewski, M.; Cao, S.; Yu, J.; Jaroniec, M. Semiconductor-based photocatalytic CO2 conversion. Mater. Horiz. 2015, 2, 261–278. [Google Scholar] [CrossRef]
  10. Maeda, K. Photocatalytic water splitting using semiconductor particles: History and recent developments. J. Photochem. Photobiol. C Photochem. Rev. 2011, 12, 237–268. [Google Scholar] [CrossRef]
  11. Arumugam, M.; Tahir, M.; Praserthdam, P. Effect of nonmetals (B, O, P, and S) doped with porous g-C3N4 for improved electron transfer towards photocatalytic CO2 reduction with water into CH4. Chemosphere 2021, 286, 131765. [Google Scholar] [CrossRef]
  12. Bi, W.; Wu, C.; Xie, Y. Atomically thin two-dimensional solids: An emerging platform for CO2 electroreduction. ACS Energy Lett. 2018, 3, 624–633. [Google Scholar] [CrossRef]
  13. Hasani, A.; Tekalgne, M.; Van Le, Q.; Jang, H.W.; Kim, S.Y. Two-dimensional materials as catalysts for solar fuels: Hydrogen evolution reaction and CO2 reduction. J. Mater. Chem. A 2019, 7, 430–454. [Google Scholar] [CrossRef]
  14. Yu, S.; Wu, X.; Wang, Y.; Guo, X.; Tong, L. 2D materials for optical modulation: Challenges and opportunities. Adv. Mater. 2017, 29, 1606128. [Google Scholar] [CrossRef]
  15. Singh, S.; Modak, A.; Pant, K.K.; Sinhamahapatra, A.; Biswas, P. MoS2–Nanosheets-based catalysts for photocatalytic CO2 reduction: A review. ACS Appl. Nano Mater. 2021, 4, 8644–8667. [Google Scholar] [CrossRef]
  16. Chhowalla, M.; Shin, H.S.; Eda, G.; Li, L.J.; Loh, K.P.; Zhang, H. The chemistry of two-dimensional layered transition metal dichalcogenide nanosheets. Nat. Chem. 2013, 5, 263–275. [Google Scholar] [CrossRef]
  17. Wang, H.; Liu, X.; Niu, P.; Wang, S.; Shi, J.; Li, L. Porous two-dimensional materials for photocatalytic and electrocatalytic applications. Matter 2020, 2, 1377–1413. [Google Scholar] [CrossRef]
  18. Jaramillo, T.F.; Jørgensen, K.P.; Bonde, J.; Nielsen, J.H.; Horch, S.; Chorkendorff, I. Identification of active edge sites for electrochemical H2 evolution from MoS2 nanocatalysts. Science 2007, 317, 100–102. [Google Scholar] [CrossRef] [Green Version]
  19. Lin, Y.K.; Chen, R.S.; Chou, T.C.; Lee, Y.H.; Chen, Y.F.; Chen, K.H.; Chen, L.C. Thickness-dependent binding energy shift in few-layer MoS2 grown by chemical vapor deposition. ACS Appl. Mater. Interfaces 2016, 8, 22637–22646. [Google Scholar] [CrossRef]
  20. Wu, J.B.; Lin, M.L.; Cong, X.; Liu, H.N.; Tan, P.H. Raman spectroscopy of graphene-based materials and its applications in related devices. Chem. Soc. Rev. 2018, 47, 1822–1873. [Google Scholar] [CrossRef] [Green Version]
  21. Ye, M.; Winslow, D.; Zhang, D.; Pandey, R.; Yap, Y.K. Recent advancement on the optical properties of two-dimensional molybdenum disulfide (MoS2) thin films. Photonics 2015, 2, 288–307. [Google Scholar] [CrossRef] [Green Version]
  22. Mignuzzi, S.; Pollard, A.J.; Bonini, N.; Brennan, B.; Gilmore, I.S.; Pimenta, M.A.; Richards, D.; Roy, D. Effect of disorder on Raman scattering of single-layer. Phys. Rev. B 2015, 91, 195411. [Google Scholar] [CrossRef] [Green Version]
  23. Feng, Z.C.; Schurman, M.; Stall, R.A. How to distinguish the Raman modes of epitaxial GaN with phonon features from sapphire substrate—Comments on “Optical properties of GaN film grown by metalorganic chemical vapor deposition” [J. Vac. Sci. Technol. A 14, 840 (1996)]. J. Vac. Sci. Technol. A Vac. Surf. Films 1997, 15, 2428–2430. [Google Scholar] [CrossRef]
  24. Mercado, E.; Goodyear, A.; Moffat, J.; Cooke, M.; Sundaram, R.S. A Raman metrology approach to quality control of 2D MoS2 film fabrication. J. Phys. D Appl. Phys. 2017, 50, 184005. [Google Scholar] [CrossRef]
  25. Zhong, W.; Deng, S.; Wang, K.; Li, G.; Li, G.; Chen, R.; Kwok, H.S. Feasible route for a large area few-layer MoS2 with magnetron sputtering. Nanomaterials 2018, 8, 590. [Google Scholar] [CrossRef] [Green Version]
  26. Lee, C.; Yan, H.; Brus, L.E.; Heinz, T.F.; Hone, J.; Ryu, S. Anomalous lattice vibrations of single-and few-layer MoS2. ACS Nano 2010, 4, 2695–2700. [Google Scholar] [CrossRef] [Green Version]
  27. Pak, S.; Lee, J.; Lee, Y.W.; Jang, A.R.; Ahn, S.; Ma, K.Y.; Cho, Y.; Hong, J.; Lee, S.; Jeong, H.Y.; et al. Strain-mediated interlayer coupling effects on the excitonic behaviors in an epitaxially grown MoS2/WS2 van der Waals heterobilayer. Nano Lett. 2017, 17, 5634–5640. [Google Scholar] [CrossRef] [Green Version]
  28. Ahn, G.H.; Amani, M.; Rasool, H.; Lien, D.H.; Mastandrea, J.P.; Ager Iii, J.W.; Dubey, M.; Chrzan, D.C.; Minor, A.M.; Javey, A. Strain-engineered growth of two-dimensional materials. Nat. Commun. 2017, 8, 608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Synnatschke, K.; Cieslik, P.A.; Harvey, A.; Castellanos-Gomez, A.; Tian, T.; Shih, C.J.; Chernikov, A.; Santos, E.J.G.; Coleman, J.N.; Claudia Backes, C. Length-and thickness-dependent optical response of liquid-exfoliated transition metal dichalcogenides. Chem. Mater. 2019, 31, 10049–10062. [Google Scholar] [CrossRef]
  30. Makuła, P.; Pacia, M.; Macyk, W. How to correctly determine the band gap energy of modified semiconductor photocatalysts based on UV–Vis spectra. J. Phys. Chem. Lett. 2018, 9, 6814–6817. [Google Scholar] [CrossRef] [Green Version]
  31. Xu, X.; Lu, R.; Zhao, X.; Xu, S.; Lei, X.; Zhang, F.; Evans, D.G. Fabrication and photocatalytic performance of a ZnxCd1− xS solid solution prepared by sulfuration of a single layered double hydroxide precursor. Appl. Catal. B Environ. 2011, 102, 147–156. [Google Scholar] [CrossRef]
  32. Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.Y.; Galli, G.; Wang, F. Emerging photoluminescence in monolayer MoS2. Nano Lett. 2010, 10, 1271–1275. [Google Scholar] [CrossRef]
  33. Eda, G.; Yamaguchi, H.; Voiry, D.; Fujita, T.; Chen, M.; Chhowalla, M. Photoluminescence from chemically exfoliated MoS2. Nano Lett. 2011, 11, 5111–5116. [Google Scholar] [CrossRef] [PubMed]
  34. Laskar, M.R.; Ma, L.; Kannappan, S.; Park, P.S.; Krishnamoorthy, S.; Nath, D.N.; Lu, W.; Wu, Y.; Rajan, S. Large area single crystal (0001) oriented MoS2. Appl. Phys. Lett. 2013, 102, 252108. [Google Scholar] [CrossRef] [Green Version]
  35. Patterson, A.L. The Scherrer formula for X-ray particle size determination. Phys. Rev. 1939, 56, 978. [Google Scholar] [CrossRef]
  36. Zhang, H.; Liu, H.; Tian, Z.; Lu, D.; Yu, Y.; Cestellos-Blanco, S.; Sakimoto, K.K.; Yang, P. Bacteria photosensitized by intracellular gold nanoclusters for solar fuel production. Nat. Nanotechnol. 2018, 13, 900–905. [Google Scholar] [CrossRef] [PubMed]
  37. Li, X.; Sun, Y.; Xu, J.; Shao, Y.; Wu, J.; Xu, X.; Pan, Y.; Ju, H.; Zhu, J.; Xie, Y. Selective visible-light-driven photocatalytic CO2 reduction to CH4 mediated by atomically thin CuIn5S8 layers. Nat. Energy 2019, 4, 690–699. [Google Scholar] [CrossRef]
  38. Linsebigler, A.L.; Lu, G.; Yates, J.T., Jr. Photocatalysis on TiO2 surfaces: Principles, mechanisms, and selected results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  39. Wu, J.; Huang, Y.; Ye, W.; Li, Y. CO2 reduction: From the electrochemical to photochemical approach. Adv. Sci. 2017, 4, 1700194. [Google Scholar] [CrossRef]
  40. Maeda, K.; Nishimura, N.; Domen, K. A precursor route to prepare tantalum (V) nitride nanoparticles with enhanced photocatalytic activity for hydrogen evolution under visible light. Appl. Catal. A Gen. 2009, 370, 88–92. [Google Scholar] [CrossRef]
  41. Jariwala, D.; Davoyan, A.R.; Tagliabue, G.; Sherrott, M.C.; Wong, J.; Atwater, H.A. Near-unity absorption in van der Waals semiconductors for ultrathin optoelectronics. Nano Lett. 2016, 16, 5482–5487. [Google Scholar] [CrossRef] [Green Version]
  42. Samaj, L. Recombination processes at grain boundaries in polycrystalline semiconductors. Phys. Status Solidi A 1987, 100, 157–167. [Google Scholar] [CrossRef]
  43. Lee, H.C.; Park, O.O. Electron scattering mechanisms in indium-tin-oxide thin films: Grain boundary and ionized impurity scattering. Vacuum 2004, 75, 275–282. [Google Scholar] [CrossRef]
  44. Godin, R.; Wang, Y.; Zwijnenburg, M.A.; Tang, J.; Durrant, J.R. Time-resolved spectroscopic investigation of charge trapping in carbon nitrides photocatalysts for hydrogen generation. J. Am. Chem. Soc. 2017, 139, 5216–5224. [Google Scholar] [CrossRef]
  45. Miao, T.J.; Tang, J. Characterization of charge carrier behavior in photocatalysis using transient absorption spectroscopy. J. Chem. Phys. 2020, 152, 194201. [Google Scholar] [CrossRef] [PubMed]
  46. Gao, Y.; Nie, W.; Wang, X.; Fan, F.; Li, C. Advanced space-and time-resolved techniques for photocatalyst studies. Chem. Commun. 2020, 56, 1007–1021. [Google Scholar] [CrossRef] [PubMed]
  47. Favaro, M.; Xiao, H.; Cheng, T.; Goddard, W.A.; Yano, J.; Crumlin, E.J. Subsurface oxide plays a critical role in CO2 activation by Cu (111) surfaces to form chemisorbed CO2, the first step in reduction of CO2. Proc. Natl. Acad. Sci. USA 2017, 114, 6706–6711. [Google Scholar] [PubMed] [Green Version]
  48. Zaera, F. New advances in the use of infrared absorption spectroscopy for the characterization of heterogeneous catalytic reactions. Chem. Soc. Rev. 2014, 43, 7624–7663. [Google Scholar] [CrossRef]
  49. Chladek, P.; Coleman, L.J.; Croiset, E.; Hudgins, R.R. Gas chromatography method for the characterization of ethanol steam reforming products. J. Chromatogr. Sci. 2007, 45, 153–157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Lum, Y.; Ager, J.W. Evidence for product-specific active sites on oxide-derived Cu catalysts for electrochemical CO2 reduction. Nat. Catal. 2019, 2, 86–93. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic diagram of the MoS2TFs synthesis process. (b) The temperature profile for the MoS2TFs growth process at different stages. (c) From left to right: the photographs of the substrate before and after MoS2TFs growth with thicknesses varying from 0 to 50 nm.
Figure 1. (a) Schematic diagram of the MoS2TFs synthesis process. (b) The temperature profile for the MoS2TFs growth process at different stages. (c) From left to right: the photographs of the substrate before and after MoS2TFs growth with thicknesses varying from 0 to 50 nm.
Catalysts 11 01295 g001
Figure 2. The AFM images (first column), height profiles (second column), and surface roughness (third column) of MoS2TFs on a sapphire substrate with different thickness: (a) 7 nm, (b) 17 nm, (c) 25 nm, and (d) 50 nm, respectively. The thickness of MoS2TFs was measured along the marked red lines.
Figure 2. The AFM images (first column), height profiles (second column), and surface roughness (third column) of MoS2TFs on a sapphire substrate with different thickness: (a) 7 nm, (b) 17 nm, (c) 25 nm, and (d) 50 nm, respectively. The thickness of MoS2TFs was measured along the marked red lines.
Catalysts 11 01295 g002
Figure 3. (a) Raman spectra of the 25 nm MoS2TFs on sapphire substrates as a function of the reaction temperature using 532 nm laser for sample excitation. The inset figure shows the Raman peak position of the LA(M) mode at 227 cm−1. (b) The LA(M) to A1g peak intensity ratio of MoS2TFs with different reaction temperatures. No clear LA(M) mode can be found in the Raman spectrum of the T = 900 °C MoS2TF sample (green color).
Figure 3. (a) Raman spectra of the 25 nm MoS2TFs on sapphire substrates as a function of the reaction temperature using 532 nm laser for sample excitation. The inset figure shows the Raman peak position of the LA(M) mode at 227 cm−1. (b) The LA(M) to A1g peak intensity ratio of MoS2TFs with different reaction temperatures. No clear LA(M) mode can be found in the Raman spectrum of the T = 900 °C MoS2TF sample (green color).
Catalysts 11 01295 g003
Figure 4. (a) Raman spectra of MoS2TFs with thickness increasing from 7 to 50 nm. (b) Raman shift of A1g and E 2 g 1 peak as a function of MoS2TFs thickness. (c) The frequency difference between the A1g and E 2 g 1 of the 16 different points on MoS2TFs as a function of the thickness. The Raman spectra of sapphire substrate, commercial monolayer, and bulk MoS2 have been used as reference.
Figure 4. (a) Raman spectra of MoS2TFs with thickness increasing from 7 to 50 nm. (b) Raman shift of A1g and E 2 g 1 peak as a function of MoS2TFs thickness. (c) The frequency difference between the A1g and E 2 g 1 of the 16 different points on MoS2TFs as a function of the thickness. The Raman spectra of sapphire substrate, commercial monolayer, and bulk MoS2 have been used as reference.
Catalysts 11 01295 g004
Figure 5. (a) UV-Vis-NIR absorption spectra of the MoS2TFs on sapphire as a function of absorbance with thickness. (b) Absorbance at the peak position of A and B excitons of the MoS2 TFs as a function of thickness. (c) Plot of A exciton energy (EA) from absorbance as a function of MoS2TFs thickness.
Figure 5. (a) UV-Vis-NIR absorption spectra of the MoS2TFs on sapphire as a function of absorbance with thickness. (b) Absorbance at the peak position of A and B excitons of the MoS2 TFs as a function of thickness. (c) Plot of A exciton energy (EA) from absorbance as a function of MoS2TFs thickness.
Catalysts 11 01295 g005
Figure 6. Tauc plots for the estimation of the band gap of MoS2TFs on sapphire with thickness increasing from (a) 7 nm, (b) 17 nm, (c) 25 nm, to (d) 50 nm, respectively. (e) The band gap of MoS2TFs as a function of thickness.
Figure 6. Tauc plots for the estimation of the band gap of MoS2TFs on sapphire with thickness increasing from (a) 7 nm, (b) 17 nm, (c) 25 nm, to (d) 50 nm, respectively. (e) The band gap of MoS2TFs as a function of thickness.
Catalysts 11 01295 g006
Figure 7. Photoluminescence spectra of (a) MoS2 thin films grown on a sapphire substrate with different thicknesses, (b) sapphire, and (c) bulk MoS2 crystal.
Figure 7. Photoluminescence spectra of (a) MoS2 thin films grown on a sapphire substrate with different thicknesses, (b) sapphire, and (c) bulk MoS2 crystal.
Catalysts 11 01295 g007
Figure 8. (a) XRD pattern of the MoS2TFs on sapphire with different thicknesses. (b) The grain size of MoS2TFs as a function of thickness.
Figure 8. (a) XRD pattern of the MoS2TFs on sapphire with different thicknesses. (b) The grain size of MoS2TFs as a function of thickness.
Catalysts 11 01295 g008
Figure 9. (a) Setup of gas-phase PC-CO2RR system. The figures below that display the PC-CO2RR activity of MoS2TFs on sapphire with different (b) reaction temperatures and (c) different thicknesses.
Figure 9. (a) Setup of gas-phase PC-CO2RR system. The figures below that display the PC-CO2RR activity of MoS2TFs on sapphire with different (b) reaction temperatures and (c) different thicknesses.
Catalysts 11 01295 g009
Figure 10. Schematic illustration of the proposed mechanism of ultra-thin film MoS2 photocatalysts for visible light-driven CO2 reduction.
Figure 10. Schematic illustration of the proposed mechanism of ultra-thin film MoS2 photocatalysts for visible light-driven CO2 reduction.
Catalysts 11 01295 g010
Table 1. The calculated grain size of MoS2TFs on sapphire from the (002) peak using Scherrer’s formula.
Table 1. The calculated grain size of MoS2TFs on sapphire from the (002) peak using Scherrer’s formula.
Thickness (nm)β/Degree2θ/DegreeD/nm
71.4314.295.60
170.6414.4712.51
250.3814.5021.07
500.3214.4725.02
Table 2. Production yields and QEs of MoS2TFs synthesized at various reaction temperatures.
Table 2. Production yields and QEs of MoS2TFs synthesized at various reaction temperatures.
Apparent Production Yield (nmol/cm2)QE (%)
MoS2TFsCH4C2H4CH3CHO C3H6O Total MoS2TFs Total (%)
Blank test 0.140.070.780.481.47Blank testnone
T = 500 °C0.830.220.960.082.09T = 500 °C0.000029
T = 700 °C0.860.391.770.143.16T = 700 °C0.000045
T = 900 °C1.380.862.350.204.79T = 900 °C0.000068
Note: The gas products from the blank test have been subtracted from all gas product values pertaining to the samples.
Table 3. Production yields and QEs of MoS2TFs with various thicknesses.
Table 3. Production yields and QEs of MoS2TFs with various thicknesses.
Apparent Production Yield (nmol/cm2)QE (%)
MoS2TFsCH4C2H4CH3CHO C3H6O Total MoS2TFs Total (%)
Blank test 0.140.070.780.481.47Blank testnone
t = 7 nm0.950.340.690.152.13t = 7 nm0.000030
t = 17 nm0.940.471.410.253.07t = 17 nm0.000044
t = 25 nm1.380.862.340.214.79t = 25 nm0.000068
t = 50 nm1.100.752.160.154.16t = 50 nm0.000059
Note: The gas products from the blank test have been subtracted from all gas product values pertaining to the samples.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Huang, Y.-F.; Liao, K.-W.; Fahmi, F.R.Z.; Modak, V.A.; Tsai, S.-H.; Ke, S.-W.; Wang, C.-H.; Chen, L.-C.; Chen, K.-H. Thickness-Dependent Photocatalysis of Ultra-Thin MoS2 Film for Visible-Light-Driven CO2 Reduction. Catalysts 2021, 11, 1295. https://doi.org/10.3390/catal11111295

AMA Style

Huang Y-F, Liao K-W, Fahmi FRZ, Modak VA, Tsai S-H, Ke S-W, Wang C-H, Chen L-C, Chen K-H. Thickness-Dependent Photocatalysis of Ultra-Thin MoS2 Film for Visible-Light-Driven CO2 Reduction. Catalysts. 2021; 11(11):1295. https://doi.org/10.3390/catal11111295

Chicago/Turabian Style

Huang, Yi-Fan, Kuan-Wei Liao, Fariz Rifqi Zul Fahmi, Varad A. Modak, Shang-Hsuan Tsai, Shang-Wei Ke, Chen-Hao Wang, Li-Chyong Chen, and Kuei-Hsien Chen. 2021. "Thickness-Dependent Photocatalysis of Ultra-Thin MoS2 Film for Visible-Light-Driven CO2 Reduction" Catalysts 11, no. 11: 1295. https://doi.org/10.3390/catal11111295

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