Surface-Induced 2D/1D Heterostructured Growth of ReS
2/CoS
2for High-Performance Electrocatalysts
Yuanwu Liu, Jing Li, Wentian Huang, Ying Zhang, Minjie Wang, Xingsen Gao, Xin Wang, Mingliang Jin, Zhipeng Hou,* Guofu Zhou, Zhang Zhang,* and Junming Liu
Cite This:ACS Appl. Mater. Interfaces2020, 12, 33586−33594 Read Online
ACCESS
Metrics & More Article Recommendations*
sı Supporting InformationABSTRACT: Two-dimensional/one-dimensional (2D/1D) het- erostructures have received much attention from researchers for their abundant catalytically active sites and low contact resistance due to formation of chemical bonds at the interface. The investigation of such heterostructures, however, is confined to lattice-matched materials, which severely limits the material candidates. Herein, we demonstrate a lattice-mismatched 2D/1D heterostructured electrocatalyst consisting of 2D ReS2nanosheets and 1D CoS2 nanowires. We propose that the higher surface energy of the CoS2nanowire and the lattice mismatch between 1D and 2D units are crucial for the growth process of ReS2nanosheets.
More importantly, the terminal S2−exposed on the surface of CoS2 nanowires serves not only as the nucleus of ReS2nanosheets but
also as a bridge to enhance electron transport efficiency. Thus, the ReS2/CoS2 heterostructures show outstanding hydrogen evolution reaction performance. This work is of general interest for the design of complex multidimensional nano-heterostructures with outstanding functionalities.
KEYWORDS: 2D/1D, electrocatalyst, ReS2, CoS2, hydrogen evolution reaction
1. INTRODUCTION
Low-dimensional nanostructured materials have found im- portant applications in catalysis because of their large specific surface area and increased structure stability.1,2 However, for single-dimensional nanostructured materials, the limited number of catalytic sites hinder further improvement of their catalytic performance. To overcome the limits on the number of catalytic sites, researchers have increasingly focused their interest on multidimensional nano-heterostructures (MDNHSs) in view of their greater catalytic activity area, which improves the catalytic performance.3−6 Moreover, MDNHSs can inherit the merits while also mitigating the drawbacks of each component element. These features are of great significance to both fundamental physics and catalytic science. As a result, MDNHSs are becoming an important area of focus for further development of catalysis.
To be an efficient catalyst, MDNHSs require not only high conductivity to realize a rapid transfer of electrons but also abundant catalytic sites.7 To date, zero-dimensional (0D) nanostructures are generally introduced to increase the number of catalytic sites in MDNHSs because their catalytic activity area is the largest among low-dimensional structures. In such 0D-based MDNHSs, however, contact resistance is relatively high because of the lack of chemical bonding and strong orbital hybridization at the interface.8−10 This feature dramatically
decreases electron transmission efficiency between materials with different numbers of dimensions, which limits further improvement of catalytic performance. It has been reported that electrons transfer faster through chemical bonds than van der Waals forces between two materials.11Thus, an ideal way to enhance electron transmission efficiency is to connect materials with different numbers of dimensions using chemical bonds at their interface. Typically, one-dimensional (1D) nanostructures possess large surface areas, and there are a large number of unsaturated bonds on the surface, which facilitate the growth of other materials.7 At the same time, 1D nanostructures have the characteristics of fast and directional electron transmission, which is conducive to the transmission of electrons to materials of other dimensions.12Hence, from the perspective of heterostructure design, the epitaxial growth of 2D materials on 1D materials is achieved by forming bridge bonds at the interface, which can effectively reduce contact resistance. Meanwhile, 2D/1D heterostructures also possess a
Received: February 16, 2020 Accepted: July 3, 2020 Published: July 3, 2020 Downloaded via NANJING UNIV on October 10, 2020 at 11:17:29 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
the performance of a mixture of the two materials. However, these heterostructures are confined to lattice-matched materials,5,14,15 which severely limits the material candidates for 2D/1D heterostructures. To extend 2D/1D systems, lattice-mismatched heterostructures must be investigated.
More importantly, recent reports suggest that a suitable lattice mismatch for heterogeneous materials may effectively improve catalytic performance by adjusting the electronic structure at the interface.16
In this work, we report the synthesis of a 2D/1D lattice- mismatched heterostructure based on a 1D CoS2nanowire and 2D ReS2 nanosheet via a chemical vapor deposition (CVD) method. X-ray photoelectron spectroscopy (XPS) results prove that the existence of terminal S2−at the 2D/1D interface plays an important role in improving electron transfer efficiency and enhancing electrical conductivity of the ReS2/CoS2. The ReS2/ CoS2 heterostructures exhibited remarkable hydrogen evolu- tion reaction (HER) performance with a small overpotential of
−114 mV at 10 mA/cm2and−196 mV at 100 mA/cm2. Our results offer new insights for both construction of 2D/1D heterostructures and synthesis of high-performance electro- catalysts in water splitting.
2. EXPERIMENTAL SECTION
2.1. Reagents.All regents were of the analytical grade and were used without further purification. Cobalt nitrate hexahydrate (Co (NO3)2·6H2O, 99%), urea (CO (NH2)2, 99%), sublimed sulfur (S, 99.99%), and sodium hypophosphite (NaH2PO2, 99%) were purchased from Aladdin. Rhenium trioxide (ReO3, 99.9%) was purchased from Alfa Aesar. Ammoniumfluoride (NH4F, 98%) was obtained from Energy Chemical. Deionized (DI) water was supplied with a Millipore system.
2.2. Synthesis of CoS2and CoP Nanowires. CoS2 nanowires were prepared according to the previous works.17,18In brief, 1.576 g of Co(NO3)2·6H2O, 1.2 g of urea, and 0.222 g of NH4F were dissolved in 35 mL of DI water and then agitated for 30 min. The obtained solution and carbon cloth (CC) (1×4 cm2) were placed into a 50 mL Teflon-lined stainless steel autoclave for hydrothermal reaction at 120°C for 6 h. After the autoclave cooled down to room temperature, the Co precursor was taken out and thoroughly washed with DI water and ethanol several times alternatively and then dried at 60 °C for 6 h. Subsequently, the Co precursor was placed in an alumina boat, and the other alumina boat containing 1 g of sulfur was placed at the upstream of the tube furnace. The two alumina boats were calcined at 500°C for 2 h with a heating speed of 2°C/min under Arflow and then cooled down to room temperature naturally.
Finally, the prepared sample was immersed in carbon disulfide (CS2) for 10 min and washed with DI water and ethanol several times. The synthetic procedures for CoP nanowires are the same as the ones for CoS2. The only difference is that the NaH2PO2 was used in the calcined step instead.
2.3. Synthesis of ReS2/CoS2and ReS2/CoP.S (500 mg) in an alumina boat was placed in the central hot zone 1 where the temperature would reach 200°C. ReO3(2 mg) in an alumina boat
was analyzed by X-ray diffraction (XRD, Bruker D8 ADVANCE) with a sweep speed of 4°min−1. Raman spectroscopy was performed on a Renishaw 42K864 system with a 532 nm excitation laser. XPS measurements were characterized using Thermo Fisher Scientific K- Alpha+.
2.5. Electrochemical Measurements.Electrochemical measure- ments were performed using an electrochemical workstation at room temperature (CHI 760C, CH Instruments Inc.) in a 0.5 M H2SO4
aqueous solution. Ag/AgCl (in a 3 M KCl solution) and a graphite rod were used as reference and counter electrodes, respectively. The prepared ReS2/CoS2heterostructure on the CC was directly used as a working electrode without any binder. Without special emphasis, the current densities were evaluated by the geometric surface area of ReS2/CoS2. All voltages were corrected to the reversed hydrogen electrode according to the following equation:ERHE=EAg/AgCl+ 0.059 pH + 0.205. The polarization curves were acquired by sweeping the potential from 0 to−0.4 V [vsreversible hydrogen electrode (RHE)]
with a scan rate of 5 mV/s. Electrochemical impedance spectroscopy (EIS) was carried out with an amplitude of 10 mV and a frequency range from 100 kHz to 0.1 Hz. Cyclic voltammetry (CV) curves at different scanning rates were used to estimate the electrochemical double layer capacitance (EDLC) of the prepared samples. Stability tests were measured by the potential between 0 and−0.4 V (vsRHE) at a sweep rate of 100 mV/s for 1000 cycles at room temperature, and linear sweep voltammetry (LSV) was recorded at a scanning rate of 5 mV/s before and after the 1000 cycles. All the electrochemical measurements have beeniR-corrected, whereiis the test current and Ris the compensation resistance.
3. RESULTS AND DISCUSSION
In the above 2D/1D lattice-mismatched heterostructure, CoS2 was chosen as the 1D component for several reasons. First, as a representative of Co-based materials and pyrite-phase tran- sition metal disulfides (TMDs), pyrite CoS2 is inherently a conductive metal, as opposed to semiconducting pyrites (NiS2, FeS2), which makes it a unique advantage as an electrocatalyst material.19Second, CoS2nanowires can expose more catalyti- cally active sites compared with bulk CoS2crystals. Third, the terminal S2−of CoS2has extremely high reactivity, which may serve as the nucleation site for lattice-mismatched materials.6,20 As an emerging 2D material, the catalytic properties of ReS2 are significantly different from those of traditional TMDs because of its weak interlayer coupling.21 Such a weak interlayer coupling feature not only allows it to expose more edge sites and basal planes, but also promote the diffusion of electrolyte ions between layers.22More importantly, S2−on the surface of CoS2 has high reactivity, which can serve as the nucleation site for other materials. For example, Zhou et al.
successfully synthesized a WS2−CoS2 electrocatalyst. The formation of WS2−CoS2is due to the high reactivity of S2−
on the surface of CoS2, which can react with (NH4)2WS4to synthesize WS2, thus forming a Co−S−W interface.23Hence, ReS2 nanosheets were expected to grow on CoS2 nanowires
because the S2−on the CoS2nanowires, with its high reactivity, should serve as a nucleation site for the growth of ReS2into a nanosheet. Moreover, both the CoS2 nanowires and ReS2 nanosheets crystallize into the cubic crystal and triclinic systems,24,25respectively, and their lattice constants have a big difference. Hence, they are ideal materials for investigating lattice-mismatched 2D/1D heterostructures in catalysis.
The ReS2(2D)/CoS2(1D) heterostructure has been synthe- sized through a two-step process, as schematically illustrated in Figure 1a−c. First, CC is chosen as the growth substrate for CoS2 nanowires. Because of the oxygen-containing functional groups on its surface, CC can bond with the Co precursor in a hydrothermal process, thus allowing the growth of CoS2 nanowire arrays. High-density ReS2 nanosheets are then grown on CoS2 nanowires via CVD because the enormous specific surface area of a nanowire provides abundant nucleation sites, and the terminal S2− in CoS2 has high reactivity.26The morphologies are investigated by SEM. CC is composed of high-density carbon fibers with an average
diameter of 7μm (Figure 1d). High-density CoS2 nanowires with lengths of 7−8 μm are uniformly distributed on the cylindrical surface of the carbon fibers (see Figures 1e and S2a). In the high-magnification inset of Figure 1e, CoS2 nanowires have a rough surface morphology, and their diameter decreases from bottom to top. As shown inFigure 1f, by using a CVD method, freestanding 2D ReS2nanosheets, which are mostly hexagonal-shaped having side lengths less than 100 nm, are synthesized on the CoS2nanowires to form a 2D/1D heterostructured system. Furthermore, high-density ReS2 nanosheets can grow uniformly on CoS2 nanowires by adjusting the growth time of ReS2nanosheets (Figure S2b,c).
Meanwhile, the thickness of a ReS2nanosheet is about 3 nm, corresponding to five layers of sandwiched S−Re−S. Clearly, compared with other low-dimensional hybrid systems, such a 2D/1D heterostructure achieves a much higher specific surface area, which could expose more catalytically active sites.27
The crystal structures are first investigated by XRD. As illustrated in Figure 2a, a characteristic peak at 14.6° Figure 1.Schematic illustration for the fabrication of heterostructured ReS2/CoS2composites. (a) Wovenfibers of bare CC. (b) CoS2nanowire arrays on CC by a hydrothermal process and sulfurization. (c) Heterostructured ReS2/CoS2on CC by CVD. Low-magnification SEM images of the (d) woven CC, (e) CoS2nanowire arrays and (f) ReS2/CoS2corresponding to (c). Insets in (d−f) are the corresponding high-magnification SEM images of a carbonfiber, a CoS2nanowire, and ReS2/CoS2, respectively.
Figure 2.(a) XRD patterns of CoS2nanowires and the ReS2/CoS2heterostructure. (b) Raman spectra of CoS2nanowires and the ReS2/CoS2 heterostructure. The migration occurs only on the Co−S bond, indicating that the growth of ReS2nanosheets affects the electron cloud of the Co− S bond. (c) XPS survey spectra of CoS2nanowires and the ReS2/CoS2heterostructure on CC. High-resolution XPS scans of (d) Co 2p, (e) Re 4f, and (f) S 2p electrons of CoS2and ReS2/CoS2, respectively.
corresponds to the (001) planes of ReS2 (JCPDS no. 52- 0818). The other peaks are highly consistent with the planes of a pyrite CoS2structure (JCPDS no. 41-1471).28Moreover, the vibration modes of molecular bonds are investigated by studying the Raman spectrum (Figure 2b). The Raman peaks at 162.5, 213.4, and 304.2 cm−1 correspond to the in- plane (Eg), out-of-plane (Ag), and like-in-plane (Eg-like) vibration modes of ReS2, respectively.29Additionally, the two peaks at 286.3 and 390.6 cm−1correspond to in-plane (Eg) and out-of-plane (Ag) vibration modes of Co−S, respectively.30 Clearly, the Raman signatures of both Egand Agmodes of the Co−S bond appearing in ReS2/CoS2 have blue shifts compared with those seen in pure CoS2. The Raman spectra indicate that the electron cloud density of S atoms on the surface of CoS2 may be influenced by the growth of ReS2. Therefore, we assume that S2−exposed on the surface of CoS2 nanowires played an important role in the initial growth of ReS2. All tests indicate that ReS2/CoS2 heterostructures are successfully prepared.
The surface electronic state and chemical composition are characterized by using XPS.Figure 2c demonstrates the whole survey spectra of the ReS2/CoS2 heterostructure and CoS2 nanowires on CC, where the characteristic peaks of Co, O, S, and C can be distinguished. In the binding energy range of 0− 500 eV, we can observe the existence of Re 4f, Re 4p, and Re 4d only in the ReS2/CoS2sample. Generally, O and C should be ascribed to the CC and the oxygen-containing functional groups attached on the surface of CC.31 In a high-resolution XPS scan of the ReS2/CoS2(seeFigure S2d), compared with the one of CC, the peak at 284.8 eV corresponds to C 1s, and the new peak appearing at 286.4 eV is ascribed to the C−S bond.32The XPS observation of the C−S bond suggests that CoS2 and CC substrates are connected by a covalent bond, which is beneficial for electron transport in CoS2/CC and for the material’s HER performance. In the high-resolution XPS spectra of the Co region (Figure 2d), the Co 2p peaks in ReS2/ CoS2 shift to a lower binding energy compared with the corresponding ones in CoS2. Such a phenomenon indicates an increase in the electron cloud density in CoS2 after
incorporating with ReS22. With regard to the Re region (Figure 2e), the peaks at 42.2 and 44.7 eV are ascribed to the binding energy of Re 4f7/2and Re 4f5/2of ReS2, respectively.33 In the S region (Figure 2f), two new peaks appear at 162.3 and 163.7 eV, which correspond to the binding energy of the Re−S bond.34 Moreover, the peaks at 162.9 and 164.1 eV are assigned to S 2p3/2and S 2p1/2of the terminal S2−in CoS2,20,35 and the two peaks shift toward lower binding energy in the ReS2/CoS2sample. We assume that the peak shift of S 2p is probably caused by the electron redistribution around S atoms at the interface of ReS2/CoS2, suggesting the existence of a coupling interface between CoS2and ReS2.36
To further investigate the interface of the ReS2/CoS2 heterostructure, scanning TEM (STEM) is used.Figure 3a is a low-magnification TEM image of the ReS2/CoS2 hetero- structure. High-density freestanding ReS2 nanosheets are distributed on CoS2 nanowires. Furthermore, as shown in the right side of Figure 3a, the high-angle annular dark-field (HAADF) TEM image and the corresponding (EDX) elemental mappings are analyzed. We notice that the Re and S elements are distributed more widely than the Co element, suggesting that high-density ReS2 nanosheets grow on the surface of the CoS2 nanowire. A small freestanding ReS2 nanosheet growing on a CoS2nanowire is displayed inFigure 3b. Clearly, the 2D crystal planes of ReS2are in a hexagonal shape. A corresponding schematic atomic diagram of the ReS2/ CoS2 heterostructure is illustrated in the right of Figure 3b.
Typically, the growth rates along different crystal planes are considered to be decisive in determining the crystal shape.37In the high-resolution STEM image of Figure 3c (inset is the enlarged view on ReS2), a distinct interface between the ReS2 nanosheet and CoS2nanowire is revealed. We confirm that Re4 units formed the 1D Re−Re chains in the ReS2nanosheet, and the two characteristic lattice spacings of 0.34 and 0.31 nm could be ascribed to (100) planes along theb-axis and (010) planes along thea-axis of ReS2, respectively.38In general, the rapid crystal growth of (100) planes is dominant, while the growth rate in the (010) direction is relatively slow.39,40Thus, the freestanding growth of ReS2 should result in anisotropic Figure 3.(a) Low-magnification TEM image of heterostructured ReS2/CoS2. The HAADF and elemental mapping images of Co, S, and Re are taken from the white boxed region. (b) HRTEM image of interfacial area on ReS2/CoS2on the left and the corresponding schematically atomic structure on the right. (c) STEM image of interfacial area between the ReS2nanosheet and CoS2nanowire. Inset is the high-magnification view of ReS2. (d) High-magnification view along the interface, with marked lattice distortion of ReS2.
2D crystal shapes. However, we observe the freestanding growth of 2D hexagonal ReS2nanosheets on CoS2nanowires, which should be related to a 2D/1D heterostructured growth mechanism. ReS2and CoS2belong to triclinic and cubic crystal systems, respectively.41,42 According to the calculation of the semiconductor heterojunction formula,43the lattice mismatch of the semiconductor heterojunction composed of these two substances is 15.22% (see Calculation of Lattice Mismatch section in theSupporting Information). During the formation of a heterostructure, the internal energy at the interface increases because of the large lattice mismatch.44As marked in Figure 3d, we observe clear lattice distortion of ReS2near the interface of ReS2/CoS2, which is a direct proof of the increased internal energy of freestanding ReS2 nanosheets grown on CoS2 nanowires. The corresponding fast Fourier transform (FFT) images of the interface region and the noninterface region are compared (see Figure S4), indicating the lattice distortion at the interface.
Typically, the terminal S2− have been exposed on the surfaces of CoS2.6,45,46 Using XPS, Han et al. demonstrated that S2−exist at the edge of CoS2and is highly catalytic.18Hou et al. confirmed the existence of terminal S2− between MoS2 and CoS2, which acted like a bridge to transfer electrons more efficiently.47In our XPS survey (Figure 2f), the terminal S2−
are also confirmed on the surfaces of CoS2 nanowires. To clarify the effect of terminal S2−on the freestanding growth of 2D hexagonal ReS2 nanosheets, cobalt phosphide (CoP) nanowires with a similar crystalline structure are selected as a control model of heterostructured growth of ReS2/CoP (see Figure S5 inSupporting Information).48,49
During the initial stage of nucleation, unsaturated S atoms on the surface of CoS2have extremely high reactivity and thus can easily react with rhenium heptoxide (Re2O7) decomposed from ReO3.50 Hence, we deduce that the terminal S2− could facilitate the nucleation of ReS2. Meanwhile, P atoms exposed on the surface of CoP are replaced by S atoms to obtain the CoPxS(1−x) structure under an S atmosphere at 500 °C.51 Theoretically, the content of S should be 0.68%, based on the content of Re (0.34%) in ReS2/CoP (seeFigure S6e). In fact, the S content is measured to be 1.39%, which suggests the existence of substituted S. The subsequent growth stage can be divided into early growth stage and final growth stage. As schematically illustrated in the growth model of Figure 4a, when 2D ReS2 grows on CoS2 with a rough surface, the internal energy is dramatically increased in the early growth stage, which inhibits planar 2D growth. On a CoP surface, however, because of a relatively smooth surface (see Figure 4b), ReS2tends to planar 2D growth. Experimentally, in the early growth stage (5 min), freestanding growth of 2D ReS2 nanosheets was observed only on CoS2nanowires (seeFigure S6a−d). Normally, the interfacial stress has an influence on the crystal morphology of 2D materials.52,53 In the final growth stage, the accumulated internal energy near the heterointerface is released, for example, the lattice distortion observed near the ReS2/CoS2heterointerface (Figure 3d). In general, the growth rate along (100) planes is higher than that along (010) and (1−10) planes. However, Hafeez et al.found that the growth of crystal planes with a higher growth rate was much inhibited at these interfaces because of the excessive internal energy.40 Thus, as schematically illustrated in thefinal stage ofFigure 4a, the growth rate in the (100) direction is somewhat reduced compared to those in (010) and (1−10) directions. Gradually, hexagonal ReS2 nanosheets appeared with increased 2D size
(seeFigure S7a). Ghoshalet al.reported that unstable regions would gradually appear in the center of planar 2D ReS2, leading to the appearance of out-of-plane unsaturated S atoms.54 The unsaturated S is reactive and can initiate freestanding 2D growth on the planar one. In fact, during the final growth stage (10 min), freestanding 2D ReS2 nanosheets also appear on the CoP nanowires and are mostly in irregular crystal shapes (see Figure S7b), which can be ascribed to the excessive growth of (100) planes.54
LSV measurements are performed to investigate the HER performances of ReS2/CoS2 electrocatalysts synthesized on CC. A typical three-electrode setup is used in 0.5 M H2SO4at a scan rate of 5 mV/s. As illustrated in Figure 5a, the HER activities with different ReS2growth times (10, 15, 20, and 25 min) are compared with bare CC, CoS2@CC, and Pt/C electrodes. The performance of the Pt/C is better than most reported ones in the literatures (Table S1). As expected, the pure CC electrode exhibits almost no HER activity, while CoS2@CC demonstrates excellent HER activity.46 The HER activities of all ReS2/CoS2@CC samples are further enhanced beyond that of CoS2@CC. The enhanced HER performance should be attributed to the S2− at the ReS2/CoS2 interface, which can improve the electron transmission efficiency at the interface. Moreover, the band diagram of ReS2/CoS2can help to understand the improvement of electronic efficiency at the interface. InFigure S8, when these two substances come into contact with each other, a Schottky junction is formed.8 Ideally,Ef‑ReS2 >Ef‑CoS2, and electrons will move from ReS2to CoS2. In the electrochemical test, a reverse bias will be applied to the electrode. Therefore, electrons willflow from CoS2to ReS2. However, the electron must cross the Schottky barrier, and the electron transmission efficiency will decrease.11As for our samples, ReS2 and CoS2 are connected by S2− on the surface of CoS2instead of simple contact. The formation of chemical bonds will change the electronic structure at the interface, thereby reducing the potential barrier.55,56Therefore, Figure 4.Schematic illustration of the three-dimensional (3D) crystal structure and growth mechanism of freestanding ReS2nanosheets. (a) In the early growth stage, ReS2 nanosheets grow out of the plane because of the rough surface of CoS2, while in thefinal growth stage, the formation of hexagonal ReS2nanosheets is attributed to interface stress. (b) On CoP, in the early growth stage, ReS2nanosheets grow parallel to the surface of CoP, while in thefinal growth stage, the formation of irregular ReS2 nanosheets should be attributed to the excess growth rate along the (100) plane.
the electron transport will be enhanced by the presence of S2−. To prove that the enhanced HER performance is not only derived from the intrinsic activity of ReS2, ReS2nanosheets are directly synthesized on CC, and their HER performance is tested (see Figure S9). In Figure S9d, the overpotentials of ReS2/CC and CoS2/CC are much smaller than that of ReS2/ CoS2at 10 mA/cm2. Compared with the HER performance of ReS2/CoS2, the performance of ReS2 or CoS2alone is poor.
Hence, the enhanced HER performance is ascribed to the introduction of ReS2 and derived from a synergistic effect between ReS2and CoS2.57Specifically, the highest-performing ReS2/CoS2-20 (seeFigure S11a) has overpotentials of 114 and 196 mV at 10 and 100 mA/cm2, respectively, which are the values closest to those of the Pt/C (28, 79 mV). Catalytically active sites of 2D ReS2existed on both the plane edges and the basal planes, which are ascribed to unsaturated S and Re−Re bonds, respectively.58−60We observe that both size and density of freestanding ReS2 nanosheets on CoS2 nanowires are gradually increased over thefirst 20 min (seeFigure S10a−d), introducing more active sites to improve HER activity.
However, with additional growth time, the HER activity of ReS2/CoS2-25 decreased sharply, corresponding to the agglomeration of high-density ReS2 nanosheets (see Figure S10e).
The Tafel slopes are analyzed byfitting the linear portions of polarization curves with the Tafel equation (η=blogj+a).61 As illustrated inFigure 5b, the Tafel slope of ReS2/CoS2-20 was 63.7 mV/dec, which is lower than those for CoS2(117.5 mV/dec), ReS2/CoS2-10 (109.5 mV/dec), ReS2/CoS2-15 (63.8 mV/dec), and ReS2/CoS2-25 (109.3 mV/dec). This slope is the closest to that of commercial Pt/C (43.17 mV/
dec). Generally, a lower Tafel slope reflects more favorable HER kinetics, suggesting that a larger current density per unit area can be obtained under a lower overpotential. To compare the electrochemical active surface area (ECSA) of electrodes, EDLC (Cdl) are analyzed by CV at different scan rates (see Figure S11b−f). As displayed in Figure 5c, the Cdl of ReS2/
CoS2-20 was 69.48 mF/cm2, which is larger than those of bare CoS2(12.24 mF/cm2), ReS2/CoS2-10 (33.94 mF/cm2), ReS2/ CoS2-15 (39.33 mF/cm2), and ReS2/CoS2-25 (32.01 mF/
cm2). ECSA can be calculated according to the Randles− Sevcik equation
C C ECSA C
40 F cm per cm
dl s
dl 2
ECSA
μ 2
= = − −
(1) whereCs(40μF/cm2) is the specific capacitance of a smooth surface.62With this equation, the ECSA of ReS2/CoS2-20 is calculated to be 1737 cm2, which is larger than those of CoS2 (306 cm2), ReS2/CoS2-10 (848 cm2), ReS2/CoS2-15 (983 cm2), and ReS2/CoS2-25 (800 cm2). From the tendency of ECSA with increased ReS2 growth time, we deduce that the agglomeration of ReS2 nanosheets can greatly reduce the number of active sites for Re−Re bonds on the plane surface.
The enhanced HER activities of ReS2/CoS2 heterostruc- tured electrocatalyst are further confirmed by EIS measure- ments at −0.2 V. Typically, Rct denotes the charge-transfer resistance for HER at the electrode/electrolyte interface. The approximate semicircle curves are presented in the Nyquist plots ofFigure 5d, where the diameter reflectsRctduring the process of HER.Rctvalues of bare CoS2,ReS2/CoS2-10, ReS2/ CoS2-15, ReS2/CoS2-20, and ReS2/CoS2-25 were 15.31, 4.86, 3.41, 2.10, and 4.85 Ω, respectively. The ReS2/CoS2-20 electrode has the lowest resistance, which suggests that it experienced faster HER kinetics than other electrodes.
Electrochemical stability is also a key factor for a high- performance electrodes. In Figure 5e, ReS2/CoS2-20 synthe- sized on CC exhibits a much improved stability at 10 mA/cm2 for a 10 h duration. The excellent stability is due to the fact that ReS2/CoS2heterostructures can prevent the oxidation of active S atoms on the surface of CoS2 nanowires. Moreover, Figure 5f demonstrates two polarization curves for ReS2/CoS2- 20 before and after a 1000-cycle stability test. The negligible change verifies that an ReS2/CoS2 heterostructure directly synthesized on CC serves as a stable and high-performance Figure 5.(a) HER polarization curves of CC, CoS2, ReS2/CoS2-10, ReS2/CoS2-15, ReS2/CoS2-20, ReS2/CoS2-25, and Pt/C electrodes in 0.5 M H2SO4with a potential sweep rate of 5 mV/s and (b) corresponding Tafel plots. (c) Dependence of current on the scan rate with different double layer capacitances. (d) Nyquist plots of CoS2, ReS2/CoS2-10, ReS2/CoS2-15, ReS2/CoS2-20, and ReS2/CoS2-25 electrodes. (e) Electrochemical stability of ReS2/CoS2-20 measured by chronopotentiometry at 10 mA/cm2. (f) HER polarization curves of ReS2/CoS2-20 before and after 1000 cycles for the stability test.
HER electrode. Based on the electrochemical performances of state-of-the-art ReS2and CoS2catalysts listed inTable S2, we have achieved the best HER performance with ReS2/CoS2. In addition, three batches of ReS2/CoS2-20 are prepared and tested for the electrochemical performance (see Figure S12), indicating a high reproducibility of the catalytic effect.
Another benchmarking methodology can be used to quickly evaluate the performance of catalysts.63 The benchmarking parameters for HER catalysts in 0.5 M H2SO4 can be all compared in one plot (seeFigure S13). The best catalyst with high catalytic activity and good stability should operate at a low potential and appear near the diagonal dashed line. Clearly, ReS2/CoS2-20 demonstrates the best HER performance among the catalysts, which is also consistent with the result shown in Figure 5a,f. In addition, we calculate its roughness factor (RF) according toeq 2.
RF A ECSA
geometric area
=
(2) Ageometric areais 1 cm2. ReS2/CoS2-20 has the largest RF (1737), indicating the largest contact area with the electrolyte.
4. CONCLUSIONS
In summary, high-density ReS2nanosheets with a controllable hexagonal 2D shape are grown freestanding on CoS2 nanowires via a CVD method, forming a hybrid 2D/1D heterostructured electrocatalyst. The corresponding growth mechanism for the 2D/1D heterostructure is proposed. We also demonstrate that S2− at the 2D/1D interface work as a bridge to enhance the electron transport efficiency of ReS2/ CoS2. Therefore, such a heterostructured electrocatalyst, with merits of both 1D and 2D components, can significantly improve HER performance. This work offers new insights for the construction of 2D/1D systems and high-performance electrocatalysts in water splitting.
■
ASSOCIATED CONTENT*sı Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsami.0c02951.
Calculation of lattice mismatch, figures showing schematic illustrations of lattice mismatch, SEM images, XPS spectrum, TEM images, STEM images, FFT images, XRD patterns, EDX elemental mapping, Raman spectrum, CV curves, HER performances of CoS2-based catalysts and Pt/C, band alignment diagram, and the benchmarking methodology (PDF)
■
AUTHOR INFORMATION Corresponding AuthorsZhipeng Hou−Guangdong Provincial Key Laboratory of Quantum Engineering and Quantum Materials, Institute for Advanced Materials, South China Academy of Advanced Optoelectronics, South China Normal University, Guangzhou 510006, P. R. China; orcid.org/0000-0003-4935-2149;
Email:[email protected]
Zhang Zhang− Guangdong Provincial Key Laboratory of Quantum Engineering and Quantum Materials, Institute for Advanced Materials, South China Academy of Advanced Optoelectronics and National Center for International Research on Green Optoelectronics, South China Academy of Advanced Optoelectronics, South China Normal University, Guangzhou
510006, P. R. China; orcid.org/0000-0001-6287-502X;
Email:[email protected] Authors
Yuanwu Liu−Guangdong Provincial Key Laboratory of Quantum Engineering and Quantum Materials, Institute for Advanced Materials, South China Academy of Advanced Optoelectronics, South China Normal University, Guangzhou 510006, P. R. China
Jing Li−Guangdong Provincial Key Laboratory of Quantum Engineering and Quantum Materials, Institute for Advanced Materials, South China Academy of Advanced Optoelectronics, South China Normal University, Guangzhou 510006, P. R.
China
Wentian Huang−Guangdong Provincial Key Laboratory of Quantum Engineering and Quantum Materials, Institute for Advanced Materials, South China Academy of Advanced Optoelectronics, South China Normal University, Guangzhou 510006, P. R. China
Ying Zhang−Guangdong Provincial Key Laboratory of Quantum Engineering and Quantum Materials, Institute for Advanced Materials, South China Academy of Advanced Optoelectronics, South China Normal University, Guangzhou 510006, P. R. China
Minjie Wang−Guangdong Provincial Key Laboratory of Quantum Engineering and Quantum Materials, Institute for Advanced Materials, South China Academy of Advanced Optoelectronics, South China Normal University, Guangzhou 510006, P. R. China
Xingsen Gao−Guangdong Provincial Key Laboratory of Quantum Engineering and Quantum Materials, Institute for Advanced Materials, South China Academy of Advanced Optoelectronics, South China Normal University, Guangzhou 510006, P. R. China; orcid.org/0000-0002-2725-0785 Xin Wang−National Center for International Research on
Green Optoelectronics, South China Academy of Advanced Optoelectronics, South China Normal University, Guangzhou 510006, P. R. China; orcid.org/0000-0002-4771-8453 Mingliang Jin−National Center for International Research on
Green Optoelectronics, South China Academy of Advanced Optoelectronics, South China Normal University, Guangzhou 510006, P. R. China
Guofu Zhou−National Center for International Research on Green Optoelectronics, South China Academy of Advanced Optoelectronics, South China Normal University, Guangzhou 510006, P. R. China; orcid.org/0000-0003-1101-1947 Junming Liu− Guangdong Provincial Key Laboratory of
Quantum Engineering and Quantum Materials, Institute for Advanced Materials, South China Academy of Advanced Optoelectronics, South China Normal University, Guangzhou 510006, P. R. China; Laboratory of Solid State Microstructure, Nanjing University, Nanjing 210093, P. R. China;
orcid.org/0000-0001-8988-8429 Complete contact information is available at:
https://pubs.acs.org/10.1021/acsami.0c02951
Author Contributions
Y.L. conducted the experiments and wrote this paper; Z.Z.
designed the experiments and guided the writing, Z.H. guided the writing; J.L., W.H., Y.Z., and M.W. helped to analyze experimental results; X.G., X.W., and M.J. offered important advices and convenient tests; Z.Z., G.Z., and J.L. funded the experiments.
■
(1) Voiry, D.; Shin, H. S.; Loh, K. P.; Chhowalla, M. Low-REFERENCES Dimensional Catalysts for Hydrogen Evolution and CO2Reduction.Nat. Rev. Chem.2018,2, 0105.
(2) Shi, Y.; Zhou, Y.; Yang, D.-R.; Xu, W.-X.; Wang, C.; Wang, F.-B.;
Xu, J.-J.; Xia, X.-H.; Chen, H.-Y. Energy Level Engineering of MoS2 by Transition-Metal Doping for Accelerating Hydrogen Evolution Reaction.J. Am. Chem. Soc.2017,139, 15479−15485.
(3) Meng, S.; Cui, Y.; Wang, H.; Zheng, X.; Fu, X.; Chen, S. Noble Metal-Free 0D-1D NiSX/CdS Nanocomposites toward Highly Efficient Photocatalytic Contamination Removal and Hydrogen Evolution under Visible Light.Dalton Trans.2018,47, 12671−12683.
(4) Gong, Y.; Zhang, X.; Yang, T.; Huang, W.; Liu, H.; Liu, H.;
Zheng, B.; Li, D.; Zhu, X.; Hu, W. Vapor Growth of CdS Nanowires/
WS2 Nanosheet Heterostructures with Sensitive Photodetections.
Nanotechnology2019,30, 345603.
(5) Li, Y.; Huang, L.; Li, B.; Wang, X.; Zhou, Z.; Li, J.; Wei, Z. Co- Nucleus 1D/2D Heterostructures with Bi2S3 Nanowire and MoS2 Monolayer: One-Step Growth and Defect-Induced Formation Mechanism.ACS Nano2016,10, 8938−8946.
(6) Yin, J.; Li, Y.; Lv, F.; Lu, M.; Sun, K.; Wang, W.; Wang, L.;
Cheng, F.; Li, Y.; Xi, P.; Guo, S. Oxygen Vacancies Dominated NiS2/ CoS2 Interface Porous Nanowires for Portable Zn−Air Batteries Driven Water Splitting Devices.Adv. Mater.2017,29, 1704681.
(7) Chen, P.; Tong, Y.; Wu, C.; Xie, Y. Surface/Interfacial Engineering of Inorganic Low-Dimensional Electrode Materials for Electrocatalysis.Acc. Chem. Res.2018,51, 2857−2866.
(8) Léonard, F.; Talin, A. A. Electrical Contacts to One- and Two- Dimensional Nanomaterials.Nat. Nanotechnol.2011,6, 773−783.
(9) Xia, F.; Perebeinos, V.; Lin, Y.-m.; Wu, Y.; Avouris, P. The Origins and Limits of Metal-Graphene Junction Resistance. Nat.
Nanotechnol.2011,6, 179−184.
(10) Smith, J. T.; Franklin, A. D.; Farmer, D. B.; Dimitrakopoulos, C. D. Reducing Contact Resistance in Graphene Devices through Contact Area Patterning.ACS Nano2013,7, 3661−3667.
(11) Wang, L.; Meric, I.; Huang, P. Y.; Gao, Q.; Gao, Y.; Tran, H.;
Taniguchi, T.; Watanabe, K.; Campos, L. M.; Muller, D. A. One- Dimensional Electrical Contact to a Two-Dimensional Material.
Science2013,342, 614−617.
(12) Zeng, Z.; Xu, R.; Zhao, H.; Zhang, H.; Liu, L.; Xu, S.; Lei, Y.
Exploration of Nanowire- and Nanotube-Based Electrocatalysts for Oxygen Reduction and Oxygen Evolution Reaction. Mater. Today Nano2018,3, 54−68.
(13) Yin, X.-L.; Li, L.-L.; Jiang, W.-J.; Zhang, Y.; Zhang, X.; Wan, L.- J.; Hu, J.-S. MoS2/CdS Nanosheets-on-Nanorod Heterostructure for Highly Efficient Photocatalytic H2 Generation under Visible Light Irradiation.ACS Appl. Mater. Interfaces2016,8, 15258−15266.
(14) Xu, B.; He, P.; Liu, H.; Wang, P.; Zhou, G.; Wang, X. A 1D/2D Helical CdS/ZnIn2S4 Nano-Heterostructure. Angew. Chem., Int. Ed.
2014,53, 2339−2343.
(15) Li, C.; Yu, Y.; Chi, M.; Cao, L. Epitaxial Nanosheet-Nanowire Heterostructures.Nano Lett.2013,13, 948−953.
(16) Sun, Y.; Xu, K.; Wei, Z.; Li, H.; Zhang, T.; Li, X.; Cai, W.; Ma, J.; Fan, H. J.; Li, Y. Strong Electronic Interaction in Dual-Cation-
136, 10053−10061.
(20) Li, F.; Wang, J.; Zheng, L.; Zhao, Y.; Huang, N.; Sun, P.; Fang, L.; Wang, L.; Sun, X. In Situ Preparation of NiS2/CoS2Composite Electrocatalytic Materials on Conductive Glass Substrates with Electronic Modulation for High-Performance Counter Electrodes of Dye-Sensitized Solar Cells.J. Power Sources2018,384, 1−9.
(21) Zang, Q.; Fu, L. Novel Insights and Perspectives into Weakly Coupled ReS2 toward Emerging Applications.Chem2019,5, 505− 525.
(22) Rahman, M.; Davey, K.; Qiao, S.-Z. Advent of 2D Rhenium Disulfide (ReS2): Fundamentals to Applications.Adv. Funct. Mater.
2017,27, 1606129.
(23) Zhou, J.; Tang, B.; Lin, J.; Lv, D.; Shi, J.; Sun, L.; Zeng, Q.; Niu, L.; Liu, F.; Wang, X.; Liu, X.; Suenaga, K.; Jin, C.; Liu, Z. Morphology Engineering in Monolayer MoS2-WS2 Lateral Heterostructures.Adv.
Funct. Mater.2018,28, 1801568.
(24) 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.
DOI: 10.1002/aenm.201870110
(25) Cui, F.; Feng, Q.; Hong, J.; Wang, R.; Bai, Y.; Li, X.; Liu, D.;
Zhou, Y.; Liang, X.; He, X.; Zhang, Z.; Liu, S.; Lei, Z.; Liu, Z.; Zhai, T.; Xu, H. Synthesis of Large-Size 1T′ReS2xSe2(1−x)Alloy Monolayer with Tunable Bandgap and Carrier Type.Adv. Mater.2017,29, 1−
705015.
(26) Liu, K.; Wang, F.; Xu, K.; Shifa, T. A.; Cheng, Z.; Zhan, X.; He, J. CoS2xSe2(1‑x)Nanowire Array: An Efficient Ternary Electrocatalyst for the Hydrogen Evolution Reaction.Nanoscale2016,8, 4699−4704.
(27) Yang, Y.; Zhang, K.; Lin, H.; Li, X.; Chan, H. C.; Yang, L.; Gao, Q. MoS2-Ni3S2 Heteronanorods as Efficient and Stable Bifunctional Electrocatalysts for Overall Water Splitting. ACS Catal. 2017, 7, 2357−2366.
(28) Guo, Y.; Gan, L.; Shang, C.; Wang, E.; Wang, J. A. Cake-Style CoS2@MoS2/RGO Hybrid Catalyst for Efficient Hydrogen Evolu- tion.Adv. Funct. Mater.2017,27, 1−602699.
(29) Keyshar, K.; Gong, Y.; Ye, G.; Brunetto, G.; Zhou, W.; Cole, D.
P.; Hackenberg, K.; He, Y.; Machado, L.; Kabbani, M.; Hart, A. H. C.;
Li, B.; Galvao, D. S.; George, A.; Vajtai, R.; Tiwary, C. S.; Ajayan, P.
M. Chemical Vapor Deposition of Monolayer Rhenium Disulfide (ReS2).Adv. Mater.2015,27, 4640−4648.
(30) Zhang, H.; Li, Y.; Xu, T.; Wang, J.; Huo, Z.; Wan, P.; Sun, X.
Amorphous Co-Doped MoS2 Nanosheet Coated Metallic CoS2 Nanocubes as an Excellent Electrocatalyst for Hydrogen Evolution.
J. Mater. Chem. A2015,3, 15020−15023.
(31) Wang, Y.; Zhu, Y.; Afshar, S.; Woo, M. W.; Tang, J.; Williams, T.; Kong, B.; Zhao, D.; Wang, H.; Selomulya, C. One-Dimensional CoS2 -MoS2 Nano-Flakes Decorated MoO2 Sub-Micro-Wires for Synergistically Enhanced Hydrogen Evolution. Nanoscale 2019, 11, 3500−3505.
(32) Yang, L.; Zhou, W.; Lu, J.; Hou, D.; Ke, Y.; Li, G.; Tang, Z.;
Kang, X.; Chen, S. Hierarchical Spheres Constructed by Defect-Rich MoS2 /Carbon Nanosheets for Efficient Electrocatalytic Hydrogen Evolution.Nano Energy2016,22, 490−498.
(33) Gao, J.; Li, L.; Tan, J.; Sun, H.; Li, B.; Idrobo, J. C.; Singh, C.
V.; Lu, T.-M.; Koratkar, N. Vertically Oriented Arrays of ReS2
Nanosheets for Electrochemical Energy Storage and Electrocatalysis.
Nano Lett.2016,16, 3780−3787.
(34) Lai, Z.; Chaturvedi, A.; Wang, Y.; Tran, T. H.; Liu, X.; Tan, C.;
Luo, Z.; Chen, B.; Huang, Y.; Nam, G.-H.; Zang, Z.; Chen, Y.; Hu, Z.;
Li, B.; Xi, S.; Zhang, Q.; Zong, Y.; Gu, L.; Kloc, C.; Du, Y.; Zhang, H.
Preparation of 1T’-Phase ReS2xSe2 ( 1‑x ) ( x = 0-1 ) Nanodots for Highly Efficient Electrocatalytic Hydrogen Evolution Reaction.J. Am.
Chem. Soc.2018,140, 8563−8568.
(35) Deng, Y. H.; Ye, C.; Tao, B. X.; Chen, G.; Zhang, Q.; Luo, H.
Q.; Li, N. B. One-Step Chemical Transformation Synthesis of CoS2 Nanosheets on Carbon Cloth as a 3D Flexible Electrode for Water Oxidation.J. Power Sources2018,397, 44−51.
(36) Wang, F.; He, P.; Li, Y.; Shifa, T. A.; Deng, Y.; Liu, K.; Wang, Q.; Wang, F.; Wen, Y.; Wang, Z.; Zhang, Z.; Sun, L.; He, J. Interface Engineered WxC@WS2 Nanostructure for Enhanced Hydrogen Evolution Catalysis.Adv. Funct. Mater.2017,27, 1−605802.
(37) Zhao, H.; Dai, Z.; Xu, X.; Pan, J.; Hu, J. Integrating Semiconducting Catalyst of ReS2 Nanosheets into P-Silicon Photo- cathode for Enhanced Solar Water Reduction. ACS Appl. Mater.
Interfaces2018,10, 23074−23080.
(38) Hämäläinen, J.; Mattinen, M.; Mizohata, K.; Meinander, K.;
Vehkamäki, M.; Räisänen, J.; Ritala, M.; Leskelä, M. Atomic Layer Deposition of Rhenium Disulfide.Adv. Mater.2018,30, 1−703622.
(39) Huang, Y.; Liu, H.; Chen, X.; Zhou, D.; Wang, C.; Du, J.; Zhou, T.; Wang, S. Density Functional Theory Investigation on Thiophene Hydrodesulfurization Mechanism Catalyzed by ReS2(001) Surface.J.
Phys. Chem. C2016,120, 12012−12021.
(40) Hafeez, M.; Gan, L.; Li, H.; Ma, Y.; Zhai, T. Large-Area Bilayer ReS2 Film/Multilayer ReS2 Flakes Synthesized by Chemical Vapor Deposition for High Performance Photodetectors.Adv. Funct. Mater.
2016,26, 4551−4560.
(41) Lin, Y.; Qiu, Z.; Li, D.; Ullah, S.; Hai, Y.; Xin, H.; Liao, W.;
Yang, B.; Fan, H.; Xu, J.; Zhu, C. NiS2@CoS2 Nanocrystals Encapsulated in N-Doped Carbon Nanocubes for High Performance Lithium/ Sodium Ion Batteries.Energy Storage Mater.2018,11, 67−
74.
(42) Zhang, Q.; Wang, W.; Zhang, J.; Zhu, X.; Zhang, Q.; Zhang, Y.;
Ren, Z.; Song, S.; Wang, J.; Ying, Z.; Wang, R.; Qiu, X.; Peng, T.; Fu, L. Highly Efficient Photocatalytic Hydrogen Evolution by ReS2via a Two-Electron Catalytic Reaction.Adv. Mater.2018,30, 1−707123.
(43) Jacques, A.; Diologent, F.; Bastie, P. In Situ Measurement of the Lattice Parameter Mismatch of a Nickel-Base Single-Crystalline Superalloy under Variable Stress.J. Mater. Sci. Eng. A2004,387−389, 944−949.
(44) Chen, J.; Wu, X.-J.; Gong, Y.; Zhu, Y.; Yang, Z.; Li, B.; Lu, Q.;
Yu, Y.; Han, S.; Zhang, Z.; Zong, Y.; Han, Y.; Gu, L.; Zhang, H. Edge Epitaxy of Two-Dimensional MoSe2and MoS2Nanosheets on One- Dimensional Nanowires.J. Am. Chem. Soc.2017,139, 8653−8660.
(45) Wang, C.; Wang, T.; Liu, J.; Zhou, Y.; Yu, D.; Cheng, J.-K.;
Han, F.; Li, Q.; Chen, J.; Huang, Y. Facile Synthesis of Silk-Cocoon S- Rich Cobalt Polysulfide as an Efficient Catalyst for the Hydrogen Evolution Reaction.Energy Environ. Sci.2018,11, 2467−2475.
(46) Zhang, J.; Liu, Y.; Sun, C.; Xi, P.; Peng, S.; Gao, D.; Xue, D.
Accelerated Hydrogen Evolution Reaction in CoS2 by Transition- Metal Doping.ACS Energy Lett.2018,3, 779−786.
(47) Hou, J.; Zhang, B.; Li, Z.; Cao, S.; Sun, Y.; Wu, Y.; Gao, Z.;
Sun, L. Vertically Aligned Oxygenated-CoS2-MoS2Heteronanosheet Architecture from Polyoxometalate for Efficient and Stable Overall Water Splitting.ACS Catal.2018,8, 4612−4621.
(48) Chang, J.; Ouyang, Y.; Ge, J.; Wang, J.; Liu, C.; Xing, W. Cobalt Phosphosulfide in Tetragonal Phase: Highly Active and Durable Catalyst for the Hydrogen Evolution Reaction. J. Mater. Chem. A 2018,6, 12353−12360.
(49) Tian, J.; Chen, J.; Liu, J.; Tian, Q.; Chen, P. Graphene Quantum Dot Engineered Nickel-Cobalt Phosphide as Highly Efficient Bifunctional Catalyst for Overall Water Splitting. Nano Energy2018,48, 284−291.
(50) Ting, L. R. L.; Deng, Y.; Ma, L.; Zhang, Y.-J.; Peterson, A. A.;
Yeo, B. S. Catalytic Activities of Sulfur Atoms in Amorphous
Molybdenum Sulfide for the Electrochemical Hydrogen Evolution Reaction.ACS Catal.2016,6, 861−867.
(51) Kibsgaard, J.; Jaramillo, T. F. Molybdenum Phosphosulfide: An Active, Acid-Stable, Earth- Abundant Catalyst for the Hydrogen Evolution Reaction**. Angew. Chem., Int. Ed. 2014, 53, 14433−
14437.
(52) Zhu, M.; Wang, J.; Holloway, B. C.; Outlaw, R. A.; Zhao, X.;
Hou, K.; Shutthanandan, V.; Manos, D. M. A Mechanism for Carbon Nanosheet Formation.Carbon2007,45, 2229−2234.
(53) Li, H.; Wu, H.; Yuan, S.; Qian, H. Synthesis and Character- ization of Vertically Standing MoS2 Nanosheets. Sci. Rep.2016, 6, 21171.DOI: 10.1038/srep21171
(54) Ghoshal, D.; Yoshimura, A.; Gupta, T.; House, A.; Basu, S.;
Chen, Y.; Wang, T.; Yang, Y.; Shou, W.; Hachtel, J. A.; et al.
Theoretical and Experimental Insight into the Mechanism for Spontaneous Vertical Growth of ReS2 Nanosheets. Adv. Funct.
Mater.2018,28, 1−801286.
(55) Zhou, Y.; Xie, Z.; Jiang, J.; Wang, J.; Song, X.; He, Q.; Ding, W.; Wei, Z. Lattice-Confined Ru Clusters with High CO Tolerance and Activity for the Hydrogen Oxidation Reaction.Nat. Catal.2020, 3, 454−462.
(56) Teitz, L.; Caspary Toroker, M. Theoretical Investigation of Dielectric Materials for Two-Dimensional Field-Effect Transistors.
Adv. Funct. Mater.2020,30, 1808544.
(57) Zhou, X.; Yang, X.; Li, H.; Hedhili, M. N.; Huang, K.-W.; Li, L.- J.; Zhang, W. Symmetric Synergy of Hybrid CoS2-WS2 Electro- catalysts for the Hydrogen Evolution Reaction. J. Mater. Chem. A 2017,5, 15552−15558.
(58) Chia, X.; Pumera, M. Layered Transition Metal Dichalcogenide Electrochemistry: Journey across the Periodic Table.Chem. Soc. Rev.
2018,47, 5602−5613.
(59) Zhou, Y.; Song, E.; Zhou, J.; Lin, J.; Ma, R.; Wang, Y.; Qiu, W.;
Shen, R.; Suenaga, K.; Liu, Q.; Wang, J.; Liu, Z.; Liu, J. Auto- Optimizing Hydrogen Evolution Catalytic Activity of ReS2 through Intrinsic Charge Engineering.ACS Nano2018,12, 4486−4493.
(60) Wang, L.; Sofer, Z.; Luxa, J.; Sedmidubský, D.; Ambrosi, A.;
Pumera, M. Layered Rhenium Sulfide on Free-Standing Three- Dimensional Electrodes is Highly Catalytic for the Hydrogen Evolution Reaction: Experimental and Theoretical Study.Electrochem.
Commun.2016,63, 39−43.
(61) Stevens, M. B.; Enman, L. J.; Batchellor, A. S.; Cosby, M. R.;
Vise, A. E.; Trang, C. D. M.; Boettcher, S. W. Measurement Techniques for the Study of Thin Film Heterogeneous Water Oxidation Electrocatalysts Measurement Techniques for the Study of Thin Film Heterogeneous Water Oxidation Electrocatalysts.Chem.
Mater.2016,29, 120−140.
(62) Chen, Y.-Y.; Zhang, Y.; Jiang, W.-J.; Zhang, X.; Dai, Z.; Wan, L.-J.; Hu, J.-S. Pomegranate-like N,P-Doped Mo2C@C Nanospheres as Highly Active Electrocatalysts for Alkaline Hydrogen Evolution.
ACS Nano2016,10, 8851−8860.
(63) McCrory, C. C. L.; Jung, S.; Ferrer, I. M.; Chatman, S. M.;
Peters, J. C.; Jaramillo, T. F. Benchmarking Hydrogen Evolving Reaction and Oxygen Evolving Reaction Electrocatalysts for Solar Water Splitting Devices.J. Am. Chem. Soc.2015,137, 4347−4357.