Next Article in Journal
Functionality and Mechanical Performance of Miniaturized Non-Assembly Pin-Joints Fabricated in Ti6Al4V by Laser Powder Bed Fusion
Next Article in Special Issue
Divalent Metal Ion Depletion from Wastewater by RVC Cathodes: A Critical Review
Previous Article in Journal
Surface Roughness-Induced Changes in Important Physical Features of CoFeSm Thin Films on Glass Substrates during Annealing
Previous Article in Special Issue
Controlling the Cooling Rate of Hydrothermal Synthesis to Enhance the Supercapacitive Properties of β-Nickel Hydroxide Electrode Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergetic Effect and Phase Engineering by Formation of Ti3C2Tx Modified 2H/1T-MoSe2 Composites for Enhanced HER

1
School of Integrated Circuits, Tsinghua University, Beijing 100084, China
2
School of Materials Science and Engineering, University of Science and Technology Beijing, Beijing 100083, China
3
Department of Physics and Tsinghua-Foxconn Nanotechnology Research Center, Tsinghua University, Beijing 100084, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(21), 6991; https://doi.org/10.3390/ma16216991
Submission received: 26 September 2023 / Revised: 26 October 2023 / Accepted: 27 October 2023 / Published: 31 October 2023
(This article belongs to the Special Issue Electrochemical Material Science and Electrode Processes)

Abstract

:
The typical semi conductivity and few active sites of hydrogen evolution of 2H MoSe2 severely restrict its electrocatalytic hydrogen evolution performance. At the same time, the 1T MoSe2 has metal conductivity and plentiful hydrogen evolution sites, making it feasible to optimize the electrocatalytic hydrogen evolution behavior of MoSe2 using phase engineering. In this study, we, through a simple one-step hydrothermal method, composed 1T/2H MoSe2, and then used newly emerging transition metal carbides with several atomic-layer thicknesses Ti3C2Tx to improve the conductivity of a MoSe2-based electrocatalyst. Finally, MoSe2@Ti3C2Tx was successfully synthesized, according to the control of the additional amount of Ti3C2Tx, to form a proper MoSe2/ Ti3C2Tx heterostructure with a better electrochemical HER performance. As obtained MoSe2@4 mg-Ti3C2Tx achieved a low overpotential, a small Tafel slope and this work offers additional insight into broadened MoSe2 and MXenes-based catalyst’s electrochemical application.

1. Introduction

Hydrogen energy is one of the most prevalent potential candidates for renewable clean energy with high energy capacity to replace fossil fuel [1,2,3]. However, up to now, the most used industrial hydrogen production is still electrochemical water splitting [4,5,6]. Electrochemical HER is a cathode double-electron-transfer reaction that begins with the Volmer step, in which electrons are transferred to the cathode surface, bind to H+ at an unoccupied active site, and produce adsorbed H*. Currently, noble metals are the most feasible catalysts for electrochemical water splitting to reduce overpotential and simultaneously facilitate reaction kinetics of hydrogen evolution reaction (HER) [7,8,9]. Their scarcity and high price significantly restrict further commercial development and large-scale applications of these noble metals for water-splitting. Hence, finding cheap and earth-abundant electrocatalytic electrodes is essential to accelerating the commercial application of hydrogen energy.
Two-dimensional transition-metal dichalcogenides (2D TMDs) have many promoting properties for HER; 2D TMD’s atomically thin-based structure increased reaction surface area and available active site [10], and different transition metals have various d orbitals. Thus, exhibiting different catalytic performance makes it feasible to improve the HER performance by doping different metal elements [11,12]. However, excellent catalytic activity and low cost compared to noble metals give it considerable potential in HER. MoSe2, as a member of the 2D transition-metal dichalcogenide (TMDC), has received widespread attention due to its corrosion stability [13], chemically tunable electronic properties [14,15,16,17,18]. MoSe2 can present as a octahedral coordination 1T (tetragonal symmetry, one layer per repeat unit, octahedral coordination) phase with metallic properties and a triangular coordination 2H (hexagonal symmetry, two lavers per repeat unit, trigonal prismatic coordination) phase with semiconductor properties. However, MoSe2 electrochemical HER performances mainly depend upon its nanostructure and Crystalline phase, but for 2H phase MoSe2, lack of active sites and weak conductivity restrict its performance in electrochemical HER [19,20,21]. Researchers use commonly used regulatory methods to improve their electrochemical performance such as strain engineering [22], surface/interface engineering [23], alloy engineering [24,25], defect engineering [26], and compounds with other materials [27,28,29,30]. However, designing and fabricating excellent MoSe2-based composite catalysts for HER is of great fundamental importance to promoting TMDC-catalyst use of clean energy.
In this study, we, through a simple one-step hydrothermal method, use an emerging type of 2D transition-metal carbide Ti3C2Tx (T = -F, -OH, -O terminal group) to improve the conductivity of MoSe2-based electrocatalysts, Ti3C2Tx. A typical MXene material is a typically a layered material with a sandwich structure consisting of three alternating layers of Ti atoms and two alternating layers of C atoms. In addition, due to the introduction of a small number of residual chemical elements during the preparation process of acid-solution etching, the surface adsorbs rich terminal groups such as -F, -OH, -O, etc. As a conductor can improve the HER performance of MoSe2 [31], the addition amount of Ti3C2Tx must be controlled accordingly to form a proper MoSe2/Ti3C2Tx heterostructure which has a better electrochemical HER performance. A small Tafel slope (79 mV dec−1) and low overpotential (−185.29 mV) were achieved with the obtained MoSe2@4 mg-Ti3C2Tx. However, the Ti3C2Tx improves the performance of 1H/2H nanoflower-like MoSe2 in HER and is still not researched. This work utilizes Ti3C2Tx as a composite substrate, and controls the MoSe2 two-phase ratio. Both of these simultaneously enhance catalytic activity, which has significance for applying TMDC materials in the HER catalytic field.

2. Experiment Section

2.1. Synthesis of Ti3C2Tx MXene Nanosheets

Firstly, via a relatively safe method, we combined a 5 wt % LiF powder and a 1 mol/HCl solution-etching in a glove box inert atmosphere, using a magnetic force-driving stirred reactor for the reaction. After heating in an oil bath for 96 h, AlF3 is generated from the bulk-layered MXene phase, and Ti3AlC2 raw material is used to remove Al and generate Ti3C2Tx. Then, ultrasonic dispersion and cleaning are carried out to obtain the required few-layer Ti3C2Tx thin sheets. The material was then dried in a vacuum-drying oven for 12 h. The product was then ground into powder in an agate mortar, collected, and vacuumed, successfully preparing layered Ti3C2Tx.

2.2. Carbon Cloth Preprocess for Hydrophilic Treatment

According to the volume ratio of 4:1, we mixed 10% HNO3 and 20% H2SO4; soaked the carbon cloth (1 cm × 1 cm) at room temperature for 12 h; took it out, and put it in ultrasonic ethanol solution for 30 min; then, used DI water wash for 30 s; and repeated 3 times to remove the residual acid. It must be sealed and stored in deionized water.

2.3. Preparation of 1T/2H Phase MoSe2@Ti3C2Tx on CC

Figure 1 shows that Se powder (1.0 mmol) was dissolved into 15% hydrazine hydrate. Magnetic solution was stirred and kept for 24 h. At the same time, a few Ti3C2Tx powders were dispersed with ultrasound (600 W) in 20 mL deionized water for 30 min; then, 1 mmol of Na2MoO4 · 2H2O was added to the dispersion and stirred for 45 min. After that, we added the reduced Se ion solution to the dispersion and sealed it in a 50 mL high-pressure reactor, using a carbon cloth (1 cm × 1 cm) as the loaded substrate, and then heated it to 180 °C for insulation for 24 h. After cooling, the sample was washed and dried in a vacuum environment of 60 °C. During the process, attention should be paid to preventing oxidation.

2.4. Characterization

The morphology of the sample was characterized using field emission scanning electron microscopy (FEI, HONG KONG, SEM, FEI Nova 400), and X-ray diffraction (XRD) patterns were recorded using Cu Ka radiation on Rigaku D/Max 4000 V to determine sample-composition information. Raman spectral information was obtained using a confocal microscope (HORIBA FRANCE LabRAM HR Evolution), using a 532 nm laser at low power. XPS (Thermo ESCALAB 250XI) determined the chemical configuration.

2.5. Electrochemical Measurement

The electrocatalytic HER measurement of the obtained products was conducted using a standard three-electrode electrochemical workstation consisting of an electrode counter electrode (Pt), a reference electrode Ag/AgCl, and a stainless-steel working electrode. All the electrocatalytic HER measurements were performed in 1 mol/L H+ solution. Electrode potential was converted into reversible hydrogen electrode potential (vs. RHE), and solution IR compensation was performed. The LSV curves of the prepared material were measured at different scanning rates in an inert constant-room-temperature environment to characterize its catalytic hydrogen evolution performance. Electrochemical impedance spectroscopy and scanning CV curves were also measured to obtain the electrochemical surface area.

3. Result and Discussion

Characterization of Ti3C2Tx

First, the Ti3C2Tx was characterized using a Scanning Electron Microscope shown in Figure 2a Ti3C2Tx which has prominent layered characteristics, which allows for a large contact area with MoSe2 and CC. However, due to introducing some terminal groups on the surface of Ti3C2 during the preparation process, the Raman spectrum was used to characterize its molecular vibration. As shown in Figure 2b, in addition to its inherent interlayer vibration mode Eg peak, there were three apparent peaks at 235 cm−1, 373 cm−1, and 610 cm−1, which, respectively, represent the vibration modes of terminal groups such as F and OH, which are the same as previously reported [32]. To determine the phase-composition structure of the Ti3C2Tx, we carried out XRD, as shown in Figure 2c. Ti3C2Tx had prominent peaks near 7.3°, 34.9°, 35.2° and 63°, which, respectively, correspond to the (002), (001), (104), (105) crystal planes of Ti3C2 [33], and the sharp peak represents its good crystallinity. However, XPS was used further to determine the element composition and valence state of Ti3C2Tx. As shown in Figure 2d, the XPS full spectra of Ti3C2Tx showed distinctive peaks of Ti and OH F, O. The Ti 2p XPS spectra after deconvoluting is shown in Figure 2e, where there are double peaks at lower binding energies of 455.6 eV and higher binding energies of 465 eV were obtained, confirming the existence of oxygen-containing terminal groups, which further confirmed our previous Raman results. The C XPS spectra after deconvoluting are shown in Figure 2f; double peaks at lower binding energies of 284.8 eV and higher binding energies of 281.8 eV were obtained after Gaussian Lorentz fitting. The stoichiometric ratio of Ti and C is about 100:68. However, The Ti3C2Tx we obtained was rich in oxygen, hydroxyl, F, and other terminal groups.
MoSe2@Ti3C2Tx was successfully synthesized on the Carbon Clothes with a facile one-step hydrothermal. During the preparation of the samples, the Ti3C2Tx played a key role, and SEM was used to observe the influence of the Ti3C2Tx on the surface morphology features of the synthesized samples. As shown in Figure 3a–c, with the increase of Ti3C2Tx content, more and more MoSe2 became attached to the carbon cloth substrate. This may be due to the intense attraction between terminal groups (-F, -OH, -O) on Ti3C2Tx and carbon fiber cloth substrate. However, the surface of MoSe2@4 Mg Ti3C2Tx had the most evenly distributed MoSe2 nanoflowers. A high-magnification scanning electron microscope was used to analyze its surface morphology further to study the MoSe2@4 Mg Ti3C2Tx. High-magnification scanning microscope photographs of MoSe2 nanoflowers are shown in Figure 3d–f. The size distribution of a single nanoflower was between 30 and 60 nm.
The Raman spectrum of all the samples show a characteristic peak at ≈285.4 cm−1. In accurate compliance to the in-plane (E12g) mode, two other peaks at ≈239.4 cm−1 and ≈342.5 cm−1 on both sides of the main peak were found, (Figure 4a). Corresponding to A1g and B12g, the characteristic Raman manifested as 2H MoSe2 [34,35,36,37]. Furthermore, there are two inconspicuous peaks at ≈194 cm−1 and ≈376 cm−1, which correspond to the characteristic Raman mode of 1T MoSe2 [38,39]. However, with the increase of Ti3C2Tx, there was no apparent shift in each peak position, and no characteristic peak of Ti3C2Tx was found in each sample; this may be due to the low amount of Ti3C2Tx or the adsorption of Ti3C2Tx on the carbon fiber cloth substrate. Raman spectra show that the MoSe2 we synthesized were 1T/2H multiple phases. To ascertain the existing phase composition and structural information of the as-synthesized samples XRD was carried out. The results are shown in Figure 4b; three prominent peaks were found which were located around 13.7°, 31.4° and 58°, respectively. Representing the (002), (110) and (008) crystal planes of MoSe2, due to the presence of surface structural defects, the (002) and (100) peaks of MoSe2 broaden in the XRD pattern [40]. Compared with the 2H-phase MoSe2 (JCPDS No 29–0914), standard XRD pattern means all the sample peaks of the (100) diffraction peaks shift right to a larger angle and the (002) crystal plane shifts left to a smaller angle, which is in good agreement with the result demonstrating the presence of 1T-phase MoSe2 in all synthesized samples [41,42]. With the increase in the amount of Ti3C2Tx, a prominent peak corresponding to the (110) crystal plane of Ti3C2 was observed at 60.5° in MoSe2@4 mg. At Ti3C2Tx, MoSe2@6 mg Ti3C2Tx, another peak was observed near 7.3° and in MoSe2@6 mg Ti3C2Tx, corresponding to the (002) crystal plane of Ti3C2. Therefore, we reached the preliminary conclusion that Ti3C2Tx did not decompose or participate in the reaction in the MoSe2@Ti3C2Tx synthesis. Next, XPS survey spectra illustrate that C, Ti, Mo, and Se elements are uniformly distributed in all samples. Next, the XPS spectra of MoSe2@Ti3C2Tx were further investigated for the presence of the 1T phase in the samples. Figure 4c shows the XPS spectrum in the Mo region. Two peaks at 54.7 and 53.7 eV correspond to 1T-phase MoSe2; double peaks at 228.3 and 231.4 eV correspond to Mo within the 1T phase, whereas 55.2 and 54.3 eV are assigned to the 2H phase MoSe2. The other pair at 229.3 eV and 232.5 eV can be assigned to the 3d5/2 and 3d3/2 orbitals of Mo within the 2H phase [43,44]. However, the 1T and 2H contents of each sample were different. The 1T-phase content in MoSe2@2 mg Ti3C2Tx was 64.4%; for MoSe2@4 mg Ti3C2Tx, it was 66.9%; and the 1T-phase content in MoSe2@6 mg Ti3C2Tx was 54.58%. However, according to the previous report, temperature plays an important role in the formation of 1T MoSe2 [45]. Theoretically, there should be little difference in the 1T MoSe2 content of the sample. However, the addition of Ti3C2Tx may have affected the synthesis of 1T-phase MoSe2, but the terminal groups of Ti3C2Tx may prevent the formation of MoSe2. All the XPS results verified that 1T and 2H phases coexist in the samples we synthesized.
We first tested the LSV curves for bare CC and MXene on CC, as shown in Figure 5. Dropping a small amount of MXene onto the carbon cloth can significantly improve the catalytic performance of the carbon cloth by about 2 mg. This may be due to the good conductivity of MXene. In addition, to determine the stability of MXene on the carbon cloth, we used a similar method in the paper 1000 CV cycles, and the performance of the sample did not change significantly. This may be due to the strong adsorption ability of MXene as a nanosheet and CC substrate, which did not considerably detach during the catalytic process. The HER electrochemical catalytic properties of MoSe2 obtained with a diverse mass of Ti3C2Tx were investigated in Figure 6a. According to XPS, the polarization curves after IR calibration are our work’s best value when the mass of Ti3C2Tx was 4 mg. The overpotential of η@10 mA/cm2 was 184. 4 mV. Still, with the increase of the Ti3C2Tx, the overpotential of @6 mg Ti3C2Tx drops to 244.8 mV. This is due to its low 1T-phase content. However, compared with the sample without Ti3C2Tx, using different masses of Ti3C2Tx suggests that the existence of Ti3C2Tx plays an important role in the majorization of HER electrocatalytic performance. The Tafel slope reflecting chemical reactivity in the HER further confirms the above results. Derived from their corresponding polarization curves, the samples prepared under different conditions are displayed in Figure 6b. The Tafel slope significantly decreased (79 mV dec−1), indicating a faster HER process in the sample. To understand the difference in the HER catalytic performance of the samples we synthesized varied Ti3C2Tx amounts in the Heyrovsky reaction rate-resolving stage. Furthermore, we analyzed CV measurement results to evaluate the effective electrochemical activity surface area (ECSA) by measuring Cdl stemming.
Samples with different component structures’ capacitive current and scan rates are presented in Figure 6c. MoSe2@4 mg-Ti3C2Tx also manifested larger Cdl (167.12 mF cm−2). Because of the existence of surface structural defects, the composite phase structure exposes more active site, which improves the performance of electrocatalytic HER. EIS tests were also conducted further to understand the electrochemical behavior in the HER process. Figure 6d presents the Nyquist plots of tested samples, displaying stable-system semicircle features. The charge-transfer resistance (Rct) can be obtained by calculating and fitting the Nyquist plot of Figure 6c. The Rct values MoSe2@4 mg-Ti3C2Tx (23 Ω), MoSe2@2 mg-Ti3C2Tx (39 Ω), MoSe2@6 mg-Ti3C2Tx (30 Ω), and MoSe2 (40 Ω) results as mentioned above expound that the addition of Ti3C2Tx is critical to enhancing HER catalytic activity. From work obtained samples, MoSe2@4 mgTi3C2Tx that exhibit the best HER electrocatalytic performance.
The double plate capacitance extracted in the nonfaradaic-zone CV cycle is proportional to the electrochemical activation area of the catalytic electrode. Figure 7 shows the cyclic voltammetry curve of the sample for different MoSe2@Ti3C2Tx. The sample was subjected to cyclic voltammetry testing. The test solution was consistent with the previous electrochemical workstation and still used a 0.5 mol/L H2SO4 solution. The scanning rates were 10, 30, 50, 70, and 90 mV/s. In addition, to ensure the stability of the sample, the last one was selected for 100 cycles at each rate and plotted. As shown in Figure 7a, the MoSe2 sample electrode without adding Ti3C2Tx showed the smallest capacitive behavior, followed by the electrode sample with the added 2 mg Ti3C2Tx. As the content of Ti3C2Tx increases, the area of the CV curve gradually increases. The sample with a dosage of 4 mg Ti3C2Tx has the largest CV area. When the dosage of Ti3C2Tx is further increased to 6 mg, the geometric area of the CV curve shows a significant reduction trend. This indicates that the doping amount of Ti3C2Tx has an optimal value. When the doping amount of Ti3C2Tx reaches an optimal value, its catalytic activity will exceed the contribution of the carbon cloth matrix, becoming the dominant role instead. However, its performance will gradually decline after exceeding a specific optimization interval. Based on the previous XPS characterization results, this is due to the addition of a large amount of Ti3C2Tx, which is affected by the adsorption of terminal groups -F, -OH, etc., on the surface of Ti3C2Tx. The solution’s environment during the preparation of 1T phase is disrupted, thereby affecting the generation of 1T-phase MoSe2, and reducing the content of 1T phase in the sample. In addition, due to the reduction of MoSe2 content in the 1T phase, the number of hydrogen evolution catalytic active sites introduced by the disordered arrangement of atoms at the interface of 1T and 2H phase in the sample due to different structures, will also be reduced. This further affects its catalytic activity and reduces the CV area at different scanning rates.
We investigated the catalytic performance of Ti3C2Tx during the in situ growth of MoSe2 electrodes on carbon cloth. The results showed that adding Ti3C2Tx optimized the electrocatalytic hydrogen evolution performance of MoSe2 electrodes to varying degrees. Based on characterization methods, it is speculated that adding Ti3C2Tx changed the in situ growth process of MoSe2 on carbon cloth. Due to the introduction of active groups, such as F during the preparation of Ti3C2Tx, it is more inclined to adhere to the carbon cloth. On the other hand, as a two-dimensional material, it is easy to synthesize heterostructures with MoSe2 based on the interlayer van der Waals force. In addition, its natural good conductivity can act as a good electronic channel between MoSe2 and the carbon cloth. Therefore, MoSe2 not only exposes more electrocatalytic active sites but also greatly improves the adhesion between the catalyst and the carbon cloth, so it can effectively optimize the electrocatalytic performance of MoSe2. Based on the successful preparation of 1T/2H impure-phase MoSe2, we investigated the effect of Ti3C2Tx addition on the electrocatalytic performance of MoSe2. For samples with a low amount of Ti3C2Tx addition, Ti3C2Tx cannot perform its expected conductivity, and the optimization of MoSe2 electrocatalytic performance is limited. However, when more Ti3C2Tx is added, the terminal groups adsorbed on the surface of Ti3C2Tx will affect the formation of 1T-phase MoSe2. With the increase of Ti3C2Tx addition, the sample with 4 mg Ti3C2Tx exhibited good electrocatalytic hydrogen evolution performance.
As shown in Table 1, the enhanced catalytic performance is attributed to a larger density of exposed catalytically active edge sites and a higher electrical conductivity. Other studies on improved catalytic activity of MoSe2 have shown that the structure of the high-content 1T phase with defects is instrumental in enhancing HER performance [46].
To study the stability of the sample, we conducted experiments on the stability of the samples MoSe2@Ti3C2Tx. After the CV test, the following experiments were carried out: the SEM of samples shown in Figure 8a, and the sample Raman spectroscopy shown in Figure 8b. There were no shifts for the Raman peak of 1T and 2H MoSe2 after the test, and the main crystal planes remain the same as before the CV test as shown in Figure 8c, and the half-peak width did not change significantly, indicating that the test did not significantly change its crystal structure. We used XPS to illustrate whether MoSe2 transitioned from 1T to 2H after the test, since phase 2H is a thermodynamically stable phase. As shown in the figure, the chemical elements and valence states that the full spectrum of XPS did not change significantly, and in spectroscopy, we used the same method as in the paper to split the Mo 3d and Se 3d peaks, respectively. However, the phase content in MoSe2@2 mg Ti3C2Tx after the CV test decreased from 64.4% to 63.3%. This reduction is tolerable since the samples were dried before testing.
In addition, the surface morphology of the samples after different cycles was observed. Figure 9a–c show the initial morphology of the sample without cycling. It can be seen that a large number of MoSe2 clusters uniformly adhere to the surface of the sample, and there is also a small amount of MoSe2 powder attached to the surface of the carbon cloth fiber. Still the contribution to hydrogen evolution performance is extremely limited due to the small contact area with the carbon cloth. A clear tissue network can be seen with significant adhesion to the substrate at high magnification. Figure 9d–f shows the surface morphology of the sample after 500 cycles, and MoSe2 powder on the surface of the carbon cloth has obviously fallen off. Still, the MoSe2 clusters adsorbed on the surface have not obviously fallen off. Still the cluster morphology has changed greatly, the MoSe2 on the surface has changed from loose nanoflower clusters to “chocolate” like whole blocks attached to the surface of the carbon cloth, and small MoSe2 particles can still be seen at local magnification. The particles are closely bound to the carbon cloth substrate. In Figure 9g–i, a large number of MoSe2 nanoparticles are still uniformly adhered to the surface of the sample after 1000 cycles. However, the difference is that the particle size is significantly reduced under local magnification, with a small amount of detachment locally. Overall, the majority of MoSe2 has a relatively tight binding force. It is worth noting that Nafion and other reagents were not used to promote the adhesion of MoSe2 on the substrate during this experimental testing process. The small amount of MoSe2 attached to the surface of the carbon cloth itself has a limited contribution to the catalytic performance of the catalyst due to its small interface with the carbon cloth, and its detachment will not significantly impact the catalyst performance.

4. Conclusions

In this work, the two-dimensional Ti3C2Tx was added to form the heterostructure with 1T/2H MoSe2 to improve the contact with the conductive carbon fiber cloth substrate. In addition, Ti3C2Tx itself, as a good conductor, can provide fast channels and more active sites for electron transfer during electrochemical reactions. According to the control, the addition of a mass of Ti3C2Tx to form proper MoSe2/Ti3C2Tx heterostructures with low overpotentials (−185.29 mV) and small Tafel slopes (79 mV dec−1) are achieved with the obtained MoSe2@4 mg-Ti3C2Tx, Moreover, phase engineering synergistic effects and remaking heterostructures using MoSe2 with other materials can also be used for MoSe2 future development in the field of electrochemical catalysis. However, how the terminal groups of Ti3C2Tx influence the formation of the 1T MoSe2 is still not clear. More comparative experiments need to be carried out with further Ti3C2Tx addition.

Author Contributions

Conceptualization, Y.W. and J.W.; Methodology, Y.W. and J.W.; Software, Q.Y.; Formal analysis, L.X.; investigation, Q.Y.; Data curation, X.Z.; Writing—original draft, L.X.; writing-review and editing, Y.W. and J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Seh, Z.W.; Kibsgaard, J.; Dickens, C.F.; Chorkendorff, I.B.; Nørskov, J.K.; Jaramillo, T.F. Combining theory and experiment in electrocatalysis: Insights into materials design. Science 2017, 355, eaad4998. [Google Scholar] [CrossRef] [PubMed]
  2. Schlapbach, L.; Züttel, A. Hydrogen-storage materials for mobile applications. Nature 2001, 414, 353–358. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2009, 8, 76–80. [Google Scholar] [CrossRef] [PubMed]
  4. Bodhankar, P.M.; Sarawade, P.B.; Kumar, P.; Vinu, A.; Kulkarni, A.P.; Lokhande, C.D.; Dhawale, D.S. Nanostructured Metal Phosphide Based Catalysts for Electrochemical Water Splitting: A Review. Small 2022, 18, 2107572. [Google Scholar] [CrossRef]
  5. Luo, Y.; Zhang, Z.; Chhowalla, M.; Liu, B. Recent Advances in Design of Electrocatalysts for High-Current-Density Water Splitting. Adv. Mater. 2021, 34, 2108133. [Google Scholar] [CrossRef]
  6. Cabán-Acevedo, M.; Stone, M.L.; Schmidt, J.R.; Thomas, J.G.; Ding, Q.; Chang, H.C.; Tsai, M.L.; He, J.H.; Jin, S. Efficient hydrogen evolution catalysis using ternary pyrite-type cobalt phosphosulphide. Nat. Mater. 2015, 14, 1245–1251. [Google Scholar] [CrossRef]
  7. Mahmood, J.; Li, F.; Jung, S.M.; Okyay, M.S.; Ahmad, I.; Kim, S.J.; Park, N.; Jeong, H.Y.; Baek, J.B. An efficient and pH-universal ruthenium-based catalyst for the hydrogen evolution reaction. Nat. Nanotechnol. 2017, 12, 441–446. [Google Scholar] [CrossRef]
  8. Han, X.; Wu, X.; Deng, Y.; Liu, J.; Lu, J.; Zhong, C.; Hu, W. Ultrafine Pt Nanoparticle-Decorated Pyrite-Type CoS2 Nanosheet Arrays Coated on Carbon Cloth as a Bifunctional Electrode for Overall Water Splitting. Adv. Energy Mater. 2018, 8, 1800935. [Google Scholar] [CrossRef]
  9. Zhao, Z.; Liu, H.; Gao, W.; Xue, W.; Liu, Z.; Huang, J.; Pan, X.; Huang, Y. Surface-Engineered PtNi-O Nanostructure with Record-High Performance for Electrocatalytic Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2018, 140, 9046–9050. [Google Scholar] [CrossRef]
  10. 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]
  11. Fu, Q.; Han, J.; Wang, X.; Xu, P.; Yao, T.; Zhong, J.; Zhong, W.; Liu, S.; Gao, T.; Zhang, Z.; et al. 2D Transition Metal Dichalcogenides: Design, Modulation, and Challenges in Electrocatalysis. Adv. Mater. 2021, 33, 1907818. [Google Scholar] [CrossRef] [PubMed]
  12. Voiry, D.; Yang, J.; Chhowalla, M. Recent Strategies for Improving the Catalytic Activity of 2D TMD Nanosheets Toward the Hydrogen Evolution Reaction. Adv. Mater. 2016, 28, 6197–6206. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, Z.; Wang, J.; Li, Y.; Zhang, S.; Xiao, L.; Wang, J.; Qi, J. Hydrothermal preparation of MoX2 (X = S, Se, Te)/TaS2 hybrid materials on carbon cloth as efficient electrocatalyst for hydrogen evolution reaction. Int. J. Hydrogen Energy 2023, 48, 4207–4219. [Google Scholar] [CrossRef]
  14. Deng, S.; Zhong, Y.; Zeng, Y.; Wang, Y.; Yao, Z.; Yang, F.; Lin, S.; Wang, X.; Lu, X.; Xia, X.; et al. Directional Construction of Vertical Nitrogen-Doped 1T-2H MoSe2/Graphene Shell/Core Nanoflake Arrays for Efficient Hydrogen Evolution Reaction. Adv. Mater. 2017, 29, 1700748. [Google Scholar] [CrossRef]
  15. Xu, C.; Peng, S.; Tan, C.; Ang, H.; Tan, H.; Zhang, H.; Yan, Q. Ultrathin S-doped MoSe2 nanosheets for efficient hydrogen evolution. J. Mater. Chem. A 2014, 2, 5597–5601. [Google Scholar] [CrossRef]
  16. Najafi, L.; Bellani, S.; Oropesa-Nuñez, R.; Ansaldo, A.; Prato, M.; Del Rio Castillo, A.E.; Bonaccorso, F. Doped-MoSe2 Nanoflakes/3d Metal Oxide–Hydr(Oxy)Oxides Hybrid Catalysts for pH-Universal Electrochemical Hydrogen Evolution Reaction. Adv. Energy Mater. 2018, 8, 1801764. [Google Scholar] [CrossRef]
  17. Chen, X.; Qiu, Y.; Liu, G.; Zheng, W.; Feng, W.; Gao, F.; Cao, W.; Fu, Y.; Hu, W.; Hu, P. Tuning electrochemical catalytic activity of defective 2D terrace MoSe2 heterogeneous catalyst via cobalt doping. J. Mater. Chem. A 2017, 5, 11357–11363. [Google Scholar] [CrossRef]
  18. Qian, J.; Wang, T.; Xia, B.; Xi, P.; Gao, D. Zn-doped MoSe2 nanosheets as high-performance electrocatalysts for hydrogen evolution reaction in acid media. Electrochim. Acta 2019, 296, 701–708. [Google Scholar] [CrossRef]
  19. Ambrosi, A.; Sofer, Z.; Pumera, M. 2H → 1T phase transition and hydrogen evolution activity of MoS2, MoSe2, WS2 and WSe2 strongly depends on the MX2 composition. Chem. Commun. 2015, 51, 8450–8453. [Google Scholar] [CrossRef]
  20. Deng, S.; Yang, F.; Zhang, Q.; Zhong, Y.; Zeng, Y.; Lin, S.; Wang, X.; Lu, X.; Wang, C.Z.; Gu, L.; et al. Phase Modulation of (1T-2H)-MoSe2/TiC-C Shell/Core Arrays via Nitrogen Doping for Highly Efficient Hydrogen Evolution Reaction. Adv. Mater. 2018, 30, 1802223. [Google Scholar] [CrossRef]
  21. Jiang, M.; Zhang, J.; Wu, M.; Jian, W.; Xue, H.; Ng, T.W.; Lee, C.S.; Xu, J. Synthesis of 1T-MoSe2 ultrathin nanosheets with an expanded interlayer spacing of 1.17 nm for efficient hydrogen evolution reaction. J. Mater. Chem. A 2016, 4, 14949–14953. [Google Scholar] [CrossRef]
  22. Wang, L.; Zeng, Z.; Gao, W.; Maxson, T.; Raciti, D.; Giroux, M.; Pan, X.; Wang, C.; Greeley, J. Tunable intrinsic strain in two-dimensional transition metal electrocatalysts. Science 2019, 363, 870–874. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, P.; Wang, B. Designing Self-Supported Electrocatalysts for Electrochemical Water Splitting: Surface/Interface Engineering toward Enhanced Electrocatalytic Performance. ACS Appl. Mater. Interfaces 2021, 13, 59593–59617. [Google Scholar] [CrossRef] [PubMed]
  24. Lin, Z.; Lin, B.; Wang, Z.; Chen, S.; Wang, C.; Dong, M.; Gao, Q.; Shao, Q.; Ding, T.; Liu, H.; et al. Facile Preparation of 1T/2H-Mo(S1−xSex)2 Nanoparticles for Boosting Hydrogen Evolution Reaction. ChemCatChem 2019, 11, 2217–2222. [Google Scholar] [CrossRef]
  25. Gong, Q.; Cheng, L.; Liu, C.; Zhang, M.; Feng, Q.; Ye, H.; Zeng, M.; Xie, L.; Liu, Z.; Li, Y. Ultrathin MoS2(1−x)Se2x Alloy Nanoflakes For Electrocatalytic Hydrogen Evolution Reaction. ACS Catal. 2015, 5, 2213–2219. [Google Scholar] [CrossRef]
  26. Kwak, I.H.; Kwon, I.S.; Lee, J.H.; Lim, Y.R.; Park, J. Chalcogen-vacancy group VI transition metal dichalcogenide nanosheets for electrochemical and photoelectrochemical hydrogen evolution. J. Mater. Chem. C 2021, 9, 101–109. [Google Scholar] [CrossRef]
  27. Zhou, X.; Liu, Y.; Ju, H.; Pan, B.; Zhu, J.; Ding, T.; Wang, C.; Yang, Q. Design and Epitaxial Growth of MoSe2–NiSe Vertical Heteronanostructures with Electronic Modulation for Enhanced Hydrogen Evolution Reaction. Chem. Mater. 2016, 28, 1838–1846. [Google Scholar] [CrossRef]
  28. Tang, X.; Zhang, J.Y.; Mei, B.; Zhang, X.; Liu, Y.; Wang, J.; Li, W. Synthesis of hollow CoSe2/MoSe2 nanospheres for efficient hydrazine-assisted hydrogen evolution. Chem. Eng. J. 2021, 404, 126529. [Google Scholar] [CrossRef]
  29. Inta, H.R.; Ghosh, S.; Mondal, A.; Tudu, G.; Koppisetti, H.V.; Mahalingam, V. Ni0.85Se/MoSe2 Interfacial Structure: An Efficient Electrocatalyst for Alkaline Hydrogen Evolution Reaction. ACS Appl. Energy Mater. 2021, 4, 2828–2837. [Google Scholar] [CrossRef]
  30. Zhang, Y.; Zhang, S.; He, Y.; Li, H.; He, T.; Shi, H.; Ma, X.; Yang, Q.; Chen, L.; Chen, J. Self-supporting MoSe2/CoSe2@CFP electrocatalyst electrode for high-efficiency HER under alkaline solution. J. Solid State Chem. 2021, 298, 122108. [Google Scholar] [CrossRef]
  31. Mahmood, A.; Khan, S.; Rahayu, F.; Shah, A.U.H.A.; Rahman, A.U.; Muhammad, T.; Khan, A.; Ullah, N. Factors Effecting and Structural Engineering of Molybdenum Nitride-Based Electrocatalyst for Overall Water Splitting: A Critical Review. Energy Technol. 2023, 11, 2194–4288. [Google Scholar] [CrossRef]
  32. Zhang, S.; Ying, H.; Huang, P.; Yang, T.; Han, W.Q. Hierarchical utilization of raw Ti3C2Tx MXene for fast preparation of various Ti3C2Tx MXene derivatives. Nano Res. 2022, 15, 2746–2755. [Google Scholar] [CrossRef]
  33. Chen, B.; Feng, A.; Liu, K.; Wu, J.; Yu, Y.; Song, L. High-performance capacitive deionization using 3D porous Ti3C2Tx with improved conductivity. J. Electroanal. Chem. 2021, 895, 115515. [Google Scholar] [CrossRef]
  34. Qi, Z.; Wang, S.; Li, Y.; Wang, L.; Zhao, L.; Ge, Q.; Zhang, J.Z. Scavenging activity and reaction mechanism of Ti3C2Tx MXene as a novel free radical scavenger. Ceram. Int. 2021, 47, 16555–16561. [Google Scholar] [CrossRef]
  35. Alhabeb, M.; Maleski, K.; Anasori, B.; Lelyukh, P.; Clark, L.; Sin, S.; Gogotsi, Y. Guidelines for Synthesis and Processing of Two-Dimensional Titanium Carbide (Ti3C2Tx MXene). Chem. Mater. 2017, 29, 7633–7644. [Google Scholar] [CrossRef]
  36. Li, C.; Zhu, L.; Li, H.; Li, H.; Wu, Z.; Liang, C.; Zhu, X.; Sun, Y. Dual surfactants applied in synthesis of MoSe2 for high-efficiency hydrogen evolution reaction. J. Alloys Compd. 2021, 863, 158092. [Google Scholar] [CrossRef]
  37. Truong, Q.D.; Nakayasu, Y.; Nguyen, Q.T.; Nguyen, D.N.; Nguyen, C.T.; Devaraju, M.K.; Rangappa, D.; Nayuki, K.; Sasaki, Y.; Tran, P.D.; et al. Defect-rich exfoliated MoSe2 nanosheets by supercritical fluid process as an attractive catalyst for hydrogen evolution in water. Appl. Surf. Sci. 2020, 505, 144537. [Google Scholar] [CrossRef]
  38. Zhao, X.; Zhao, Y.; Huang, B.; Cai, W.; Sui, J.; Yang, Z.; Wang, H.E. MoSe2 nanoplatelets with enriched active edge sites for superior sodium-ion storage and enhanced alkaline hydrogen evolution activity. Chem. Eng. J. 2020, 382, 123047. [Google Scholar] [CrossRef]
  39. Ullah, N.; Xie, M.; Hussain, S.; Yaseen, W.; Shah, S.A.; Yusuf, B.A.; Oluigbo, C.J.; Rasheed, H.U.; Xu, Y.; Xie, J. Simultaneous synthesis of bimetallic@3D graphene electrocatalyst for HER and OER. Front. Mater. Sci. 2021, 15, 305–315. [Google Scholar] [CrossRef]
  40. Setayeshgar, S.; Karimipour, M.; Molaei, M.; Moghadam, M.R.; Khazraei, S. Synthesis of scalable 1T/2H–MoSe2 nanosheets with a new source of Se in basic media and study of their HER activity. Int. J. Hydrogen Energy 2020, 45, 6090–6101. [Google Scholar] [CrossRef]
  41. Yin, Y.; Zhang, Y.; Gao, T.; Yao, T.; Zhang, X.; Han, J.; Wang, X.; Zhang, Z.; Xu, P.; Zhang, P.; et al. Synergistic Phase and Disorder Engineering in 1T-MoSe2 Nanosheets for Enhanced Hydrogen-Evolution Reaction. Adv. Mater. 2017, 29, 1700311. [Google Scholar] [CrossRef] [PubMed]
  42. Shi, H.; Zhang, H.; Li, M.; Wang, Y.; Wang, D. Nanoflower-like 1T/2H mixed-phase MoSe2 as an efficient electrocatalyst for hydrogen evolution. J. Alloys Compd. 2021, 878, 160381. [Google Scholar] [CrossRef]
  43. Zhang, J.; Chen, Y.; Liu, M.; Du, K.; Zhou, Y.; Li, Y.; Wang, Z.; Zhang, J. 1T@2H-MoSe2 nanosheets directly arrayed on Ti plate: An efficient electrocatalytic electrode for hydrogen evolution reaction. Nano Res. 2018, 11, 4587–4598. [Google Scholar] [CrossRef]
  44. Li, N.; Wu, J.; Lu, Y.; Zhao, Z.; Zhang, H.; Li, X.; Zheng, Y.Z.; Tao, X. Stable multiphasic 1T/2H MoSe2 nanosheets integrated with 1D sulfide semiconductor for drastically enhanced visible-light photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2018, 238, 27–37. [Google Scholar] [CrossRef]
  45. Li, N.; Zhang, Y.; Jia, M.; Lv, X.; Li, X.; Li, R.; Ding, X.; Zheng, Y.Z.; Tao, X. 1T/2H MoSe2-on-MXene heterostructure as bifunctional electrocatalyst for efficient overall water splitting. Electrochim. Acta 2019, 326, 134976. [Google Scholar] [CrossRef]
  46. Zhang, L.; Sun, L.; Huang, Y.; Sun, Y.; Hu, T.; Xu, K.; Ma, F. Hydrothermal synthesis of N-doped RGO/MoSe2 composites and enhanced electro-catalytic hydrogen evolution. J. Mater. Sci. 2017, 52, 13561–13571. [Google Scholar] [CrossRef]
  47. Vikraman, D.; Hussain, S.; Prasanna, K.; Karuppasamy, K.; Jung, J.; Kim, H.S. Facile method to synthesis hybrid phase 1T@2H MoSe2 nanostructures for rechargeable lithium ion batteries. J. Electroanal. Chem. 2019, 833, 333–339. [Google Scholar] [CrossRef]
  48. Mao, Z.; Wang, C.; Lu, H.; Tang, K.; Li, Q.; Yan, C.; Wang, X. Boron-modified electron transfer in metallic 1T MoSe2 for enhanced inherent activity on per-catalytic site toward hydrogen evolution. Adv. Mater. Interfaces 2020, 7, 1901560. [Google Scholar] [CrossRef]
  49. Maity, S.; Das, B.; Samanta, M.; Das, B.K.; Ghosh, S.; Chattopadhyay, K.K. MoSe2-amorphous CNT hierarchical hybrid core–shell structure for efficient hydrogen evolution reaction. ACS Appl. Energ. Mater. 2020, 3, 5067–5076. [Google Scholar] [CrossRef]
  50. Ren, X.; Yao, Y.; Ren, P.; Wang, Y.; Peng, Y. Facile sol-gel synthesis of C@MoSe2 coreshell composites as advanced hydrogen evolution reaction catalyst. Mater. Lett. 2019, 238, 286–289. [Google Scholar] [CrossRef]
  51. Shi, Y.; Gao, W.; Lu, H.; Huang, Y.; Zuo, L.; Fan, W.; Liu, T. Carbon-nanotube-incorporated graphene scroll-sheet conjoined aerogels for efficient hydrogen evolution reaction. ACS Sustain. Chem. Eng. 2017, 5, 6994–7002. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of hydrothermal preparation of 1T/2H MoSe2@Ti3C2Tx.
Figure 1. Schematic diagram of hydrothermal preparation of 1T/2H MoSe2@Ti3C2Tx.
Materials 16 06991 g001
Figure 2. (a) SEM image; (b) Raman Spectrum; (c) XRD; and (df) XPS of Ti3C2Tx.
Figure 2. (a) SEM image; (b) Raman Spectrum; (c) XRD; and (df) XPS of Ti3C2Tx.
Materials 16 06991 g002
Figure 3. SEM images of synthesized MoSe2@Ti3C2Tx/CC samples; (a) MoSe2@2 mgTi3C2Tx/CC; (b) MoSe2@4 mgTi3C2Tx/CC; (c) MoSe2@6 mgTi3C2Tx/CC; and MoSe2 nanoflower on the samples of MoSe2@4 mgTi3C2Tx/CC at (d) ×5000, (e) ×100,000, (f) ×200,000.
Figure 3. SEM images of synthesized MoSe2@Ti3C2Tx/CC samples; (a) MoSe2@2 mgTi3C2Tx/CC; (b) MoSe2@4 mgTi3C2Tx/CC; (c) MoSe2@6 mgTi3C2Tx/CC; and MoSe2 nanoflower on the samples of MoSe2@4 mgTi3C2Tx/CC at (d) ×5000, (e) ×100,000, (f) ×200,000.
Materials 16 06991 g003
Figure 4. MoSe2 synthesized with different Ti3C2Tx-added amounts; (a) Raman spectrometer shifts; (b) XRD pattern; (c) Mo 3d XPS spectra; and (d) Se 3d XPS spectra.
Figure 4. MoSe2 synthesized with different Ti3C2Tx-added amounts; (a) Raman spectrometer shifts; (b) XRD pattern; (c) Mo 3d XPS spectra; and (d) Se 3d XPS spectra.
Materials 16 06991 g004
Figure 5. (a) LSV of Bare CC and MXene on CC; (b) LSV of MXene on CC and MXene on CC after 1000 CV cycles.
Figure 5. (a) LSV of Bare CC and MXene on CC; (b) LSV of MXene on CC and MXene on CC after 1000 CV cycles.
Materials 16 06991 g005
Figure 6. Electrochemical properties of MoSe2@ Ti3C2Tx synthesized with diverse amount of Ti3C2Tx. (a) Polarization curves after IR calibration; (b) corresponding Tafel plots of the samples stemming from (a); (c) Nyquist plots of the samples; and (d) Cdl current at different scan rates.
Figure 6. Electrochemical properties of MoSe2@ Ti3C2Tx synthesized with diverse amount of Ti3C2Tx. (a) Polarization curves after IR calibration; (b) corresponding Tafel plots of the samples stemming from (a); (c) Nyquist plots of the samples; and (d) Cdl current at different scan rates.
Materials 16 06991 g006
Figure 7. CV scanning curve in the scanning range of 0.1 V–0.2 V for different samples of (a) MoSe2, (b) MoSe2@2 mg-Ti3C2Tx, (c) MoSe2@4 mg-Ti3C2Tx, and (d) MoSe2@6 mg-Ti3C2Tx.
Figure 7. CV scanning curve in the scanning range of 0.1 V–0.2 V for different samples of (a) MoSe2, (b) MoSe2@2 mg-Ti3C2Tx, (c) MoSe2@4 mg-Ti3C2Tx, and (d) MoSe2@6 mg-Ti3C2Tx.
Materials 16 06991 g007
Figure 8. Characterization of MoSe2@2 mg Ti3C2Tx after CV test samples of (a) SEM, (b) Raman spectrum, (c) XRD spectrum, (d) XPS full spectrum, (e) XPS Mo 3d peak, and (f) XPS Se 3d peak.
Figure 8. Characterization of MoSe2@2 mg Ti3C2Tx after CV test samples of (a) SEM, (b) Raman spectrum, (c) XRD spectrum, (d) XPS full spectrum, (e) XPS Mo 3d peak, and (f) XPS Se 3d peak.
Materials 16 06991 g008
Figure 9. (ac) Carbon cloth in situ growth MoSe2@4 mg Ti3C2Tx initial surface morphology of sample (df) after 500 Cycles, and microscopic morphology of the sample at different magnification times, and (gi) after 1000 cycles, and microscopic morphology of the sample at different magnification times.
Figure 9. (ac) Carbon cloth in situ growth MoSe2@4 mg Ti3C2Tx initial surface morphology of sample (df) after 500 Cycles, and microscopic morphology of the sample at different magnification times, and (gi) after 1000 cycles, and microscopic morphology of the sample at different magnification times.
Materials 16 06991 g009
Table 1. Electrochemical performance (j = 10 mA cm−2) comparison of different MoSe2-based composite samples.
Table 1. Electrochemical performance (j = 10 mA cm−2) comparison of different MoSe2-based composite samples.
SampleOverpotential
(V vs. RHE)
Tafel Slope
(mV dec−1)
References
2H MoSe20.30082[47]
MoSe2-amorphous CNT0.25449[48]
C@MoSe20.27072[49]
Graphene-carbon nanotube aerogel-MoSe2 hybrid0.22868[50]
N-doped RGO/MoSe2 composites0.22978[51]
MoSe2/Ti3C2Tx0.18579This work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xiao, L.; Yang, Q.; Zhu, X.; Wei, Y.; Wang, J. Synergetic Effect and Phase Engineering by Formation of Ti3C2Tx Modified 2H/1T-MoSe2 Composites for Enhanced HER. Materials 2023, 16, 6991. https://doi.org/10.3390/ma16216991

AMA Style

Xiao L, Yang Q, Zhu X, Wei Y, Wang J. Synergetic Effect and Phase Engineering by Formation of Ti3C2Tx Modified 2H/1T-MoSe2 Composites for Enhanced HER. Materials. 2023; 16(21):6991. https://doi.org/10.3390/ma16216991

Chicago/Turabian Style

Xiao, Lei, Qichao Yang, Xiangyang Zhu, Yang Wei, and Jing Wang. 2023. "Synergetic Effect and Phase Engineering by Formation of Ti3C2Tx Modified 2H/1T-MoSe2 Composites for Enhanced HER" Materials 16, no. 21: 6991. https://doi.org/10.3390/ma16216991

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