Huahao Gua,
Longsheng Zhanga,
Yunpeng Huanga,
Youfang Zhanga,
Wei Fan*b and
Tianxi Liu*ab
aState Key Laboratory of Molecular Engineering of Polymers, Department of Macromolecular Science, Fudan University, 220 Handan Road, Shanghai 200433, P. R. China. E-mail: txliu@fudan.edu.cn; Fax: +86-21-65640293; Tel: +86-21-55664197
bState Key Laboratory of Modification of Chemical Fibers and Polymer Materials, College of Materials Science and Engineering, Donghua University, 2999 North Renmin Road, Shanghai 201620, P. R. China. E-mail: 10110440003@fudan.edu.cn
First published on 25th January 2016
Electrolysis of water is a sustainable and environmentally friendly way to produce hydrogen, which has motivated people to develop efficient and earth-abundant electrocatalysts that minimize energy consumption. Herein, graphene nanoribbon@MoS2 (GNR@MoS2) hybrids with hierarchical structure have been facilely fabricated as efficient electrocatalysts for the hydrogen evolution reaction (HER). Derived from longitudinally unzipping of multi-walled carbon nanotubes, GNR sheets can provide a greater surface area for the decoration of MoS2, which not only stems from the outer wall sheets, but also from the additional exfoliated inner wall space, as well as from the unique ribbon edges. Furthermore, the interconnected GNR sheets can form a conductive pathway for fast electron transportation and an open structure for convenient electrolyte permeation. As a consequence, the GNR@MoS2 hybrids exhibit excellent electrochemical activity as HER catalysts with a low onset potential of −0.11 V vs. the reversible hydrogen electrode and a small Tafel slope of 43.4 mV per decade. The outstanding electrocatalytic performance of the GNR@MoS2 hybrids can be ascribed to their unique hierarchical architecture with numerous active sites, as well as synergistic effects between the electrocatalytic MoS2 nanosheets and conductive GNR framework, making them promising materials for future electrocatalysts in the HER.
Recent studies have confirmed that transition metal chalcogenides, such as molybdenum and tungsten sulfide (MoSx and WS2, x ≈ 2 to 3), will become promising candidates in the application of HER electrocatalysts.10,11 MoS2 possesses a lamellar hexagonal structure, in which the intra-layer Mo–S bonds are covalently bonded while the adjacent layers are weakly interacting by van der Waals forces.12,13 It is reported that the hydrogen binding energy of MoS2 is close to that of Pt, implying that MoS2 can reduce protons at low overpotentials.14,15 Furthermore, both experimental16 and computational14 studies have concluded that HER activity arises from the sites located along the edges of MoS2 layers, while the basal surfaces are catalytically inert. As a result, emerging studies have sought to modulate the nanostructures of MoS2 to maximally increase its exposure of active edge sites for catalysis.17–20 For example, Jaramillo et al. designed a fully contiguous 3D mesoporous MoS2 thin film, in which the double-gyroid structure can preferentially expose more catalytically active edge sites.21 A scalable method was provided by Xie et al. to incorporate defects into MoS2 surfaces, which led to partial cracking of the catalytically inert basal planes, and at the same time, engendered additional density of exposed active sites, thus resulting in a small onset potential of −120 mV with a Tafel slope of 50 mV per decade.22
Although increasing the number of active edge sites can improve the electrochemical performance of MoS2, the poor intrinsic conductivity and anisotropic electrical transport of MoS2 significantly impede its overall catalytic performance. Therefore, the key challenge to apply MoS2 in the HER lies in the compensation for the conductivity while maintaining its nanosize. In this regard, forming highly dispersed and nanosized (i.e., edge rich) MoS2 on a conducting matrix is an ideal protocol to improve its catalytic activity for the HER. As a result, various conductive substrates, such as reduced graphene oxide,23 carbon nanotubes (CNTs),24 and 3D graphene foam25,26 are chosen for mediating the growth of MoS2, which not only efficiently prevents the aggregation of MoS2, but also tremendously improves the conductivity of the whole composite material.
Among the various carbon materials, CNTs have been extensively studied in a wide range of applications due to their high electrical conductivity and good mechanical properties. However, for the structure with both ends closed, the inner surface and intra tube space of CNTs are chemically inert, which greatly restricts their potential in the energy conversion and storage field.27–29 Consequently, it is meaningful to open the relatively closed structure of CNTs, especially for multi-walled CNTs (MWCNTs). Through longitudinally unzipping single or multi-walled CNTs, the tubular structure of CNTs can be opened up, thus obtaining graphene nanoribbons (GNRs). GNRs are quasi-one-dimensional, narrow and flat stripes of graphene, with the size confinement in two coordinates. Quasi-one-dimensional structures typically refer to some morphology, including nanowires, core–shell nanowires, nanotubes, nanobelts, nanoribbons, nanorods and nanorings.30 Thanks to this unique geometry, the most striking characteristics of GNRs are their increased surface area and the abundant existence of exfoliated ribbon edge planes. As a kind of novel carbon nanomaterial, quasi-one-dimensional GNRs not only inherit many outstanding characteristics of traditional carbon materials, such as lightweight, high thermal conductivity, and excellent electrical conductivity, but also possess various interesting properties due to their distinctive edge sites, such as edge dependent electronic properties and variation in the band gap owing to electron confinement, which are the significant differences between CNTs and graphene.31 Although there are many methods which can produce a microscopic specimen of GNRs, such as chemical vapor deposition (CVD),32 and lithographic33 and plasma etching,34 they not only require precise equipment or harsh experimental conditions, but also cannot be used to produce GNRs with high yields and controlled widths.35 Nevertheless, the solution-based oxidative procedure via longitudinal unzipping of MWCNTs followed by reduction is a simple, efficient, and scalable method to prepare GNRs, which is reported by Tour and co-workers.36 The produced GNRs by this method possess excellent surface integration with fewer defects or holes on the basal plane and a high length-to-width ratio. Moreover, because of their unique architecture, quasi-one-dimensional GNRs also exhibit some novel capabilities in electrochemistry. For instance, GNRs exhibit enhanced electrochemical lithium storage performance as compared with other conventionally used carbon materials, which is attributed to the existence of a large number of edge sites.37 Yi et al. prepared nitrogen-doped GNR aerogels as efficient electrocatalysts for the oxygen reduction reaction, in which GNRs make full use of their high electrical conductivity and endow extra catalytic sites without destroying the structural integrity.38 A vertical MWCNT carpet was split to form a GNR carpet for application in supercapacitors. The splitting treatment increased the specific capacitance by about 4 times, which was largely induced by the enlarged specific surface area in the additional intra tube space and inner tube surface.39 Therefore, these distinctive features of GNRs have made them promising candidates in electrochemical fields.
In this work, GNRs are employed as conductive substrates for the growth of few-layered MoS2 nanosheets via a facile solvothermal method. In this hybrid, the increased surface space along with unzipped ribbon edges of the GNR sheets provide more sites for the anchoring of MoS2, which greatly enlarges the exposed catalytic edge sites. Moreover, the intimate contact between the MoS2 nanosheets and underlying GNR network further promotes fast electron transport from the less conductive MoS2 nanosheets to the electrodes. Last but not least, the unzipping of MWCNTs into GNR sheets can benefit the permeation of the electrolyte, as the thinnish ribbon platelets make it easier for ion diffusion compared with the robust walls of MWCNTs. Therefore, the rationally designed GNR@MoS2 hybrids exhibit excellent HER performance with a low onset potential of −0.11 V, large cathodic currents, as well as a small Tafel slope of 43.4 mV per decade. Besides preparing a GNR-based hybrid with a hierarchical structure as an efficient Pt-free catalyst in the HER, this report also extends the application of GNRs, making them a promising substrate in the electrochemical field.
All electrochemical tests were performed in a standard three-electrode setup using a CHI 660D electrochemical workstation (Shanghai Chenhua Instrument Co., China), where a graphite rod was used as the counter electrode, a saturated calomel electrode (SCE) as the reference electrode and the modified GCE as the working electrode. Linear sweep voltammetry (LSV) was measured with a scan rate of 5 mV s−1 in 0.5 M H2SO4. All polarization curves were corrected for IR loss and all of the potentials were calibrated to a reversible hydrogen electrode (RHE). The cycling stability was investigated by cyclic voltammetry (CV) between −0.35 and 0.25 V vs. RHE at a scan rate of 100 mV s−1. AC impedance measurements were carried out with frequencies ranging from 106 Hz to 10−2 Hz with an amplitude of 5 mV.
In the solvothermal process, MoS2 nanosheets are decorated onto the surface of the GONR sheets with different loading amounts while the GONRs are reduced to GNRs by N2H4 at the same time. The morphology of the GNR@MoS2 hybrids is shown in Fig. 2. It can be seen in Fig. 2A and B that ultrathin MoS2 nanosheets are decorated onto the GNRs, forming the crumpled edges in the GNR@MoS2-0.5 hybrid. However, there are still some exposed GNRs not fully covered by the MoS2 nanosheets. For GNR@MoS2-1, the layered MoS2 nanosheets are uniformly distributed on the GNR surface, forming a three-dimensional (3D) network (Fig. 2C and D). The unique structure of the quasi-one-dimensional GNRs with unzipped ribbon edges can provide more active sites for the growth of MoS2, thus enhancing the electrocatalytic performance. The EDS mapping images of GNR@MoS2-1 are shown in Fig. S2,† which confirms that the MoS2 nanosheets are distributed evenly on the GNRs and the molar ratio of S to Mo is about 2:1. With the increase of the amount of (NH4)2MoS4 precursor (Fig. 2E and F), the MoS2 nanosheet layers tend to become thicker and some aggregations appear on the surface of the GNRs in the GNR@MoS2-2 hybrid. For pure MoS2 without the GNR template (Fig. S3A†), it can be observed that MoS2 seriously aggregates into a hydrangea-like morphology, which greatly hinders the exposure of the catalytically active edge sites of MoS2. Fig. S3B† displays the morphology of the GNR sheets produced in the same solvothermal reaction without the addition of (NH4)2MoS4. After reduction with N2H4, the GNR sheets become a little frizzy, which may result from the removal of oxygen-containing groups.
Fig. 2 FESEM images of the (A and B) GNR@MoS2-0.5, (C and D) GNR@MoS2-1, and (E and F) GNR@MoS2-2 hybrids at low (left) and high (right) magnifications. |
The detailed microstructure of the GNR@MoS2-1 hybrid was further confirmed by TEM observations, as shown in Fig. 3. The surface of the GNR sheets is fully covered with curly ultrathin MoS2 nanosheets, which is in good accordance with the FESEM observations (Fig. 2C and D). In addition, it can be obviously found that the fully exfoliated ribbon edges are also anchored with some corrugated MoS2 nanosheets (white arrows in Fig. 3A), extending from the strip-like basal plane, which offers additional active sites for the growth of MoS2, making full utilization of the unique quasi-one-dimensional structure of the GNR nanosheets. The HRTEM image in Fig. 3B further elucidates that the MoS2 nanosheets are composed of 5–7 sandwiched S–Mo–S layers, along with an inter-layer spacing of 0.69 nm. In order to make a more distinct comparison, the morphology of the MWCNT@MoS2 hybrid is displayed in Fig. S4.† Comparing the two images, it can be obviously observed that the average width of GNR@MoS2-1 is larger than that of MWCNT@MoS2, which is generated by fully unzipping the tube walls of the MWCNTs. As a result, more active sites, both the outer and inner wall surfaces, are presented in the GNR sheets for the decoration of the MoS2 sheets as compared with that of the MWCNTs. Naturally, with the same amount of (NH4)2MoS4 precursor, the MoS2 sheets grow more densely and closely on the substrate of the quasi-one-dimensional GNR sheets with a large number of extra growth sites originating from the unzipping process. On the contrary, the MWCNTs only provide the space of the outer wall for the decoration of MoS2 sheets, which can be verified by the TEM image, where the trace of the inner wall boundaries could still be slightly tracked even after the growth of the MoS2 sheets (white arrows in Fig. S4†).
XRD patterns of the GNRs, pure MoS2 and the GNR@MoS2 hybrids are displayed in Fig. 4. Different from the GONRs showing a diffraction peak at 2θ = 9.2° in Fig. S1,† the pattern of the GNRs displays a broad peak at 2θ = 25.8°, indicating that the inter-layer space of the GNRs is decreased due to the removal of oxygen-containing groups. A similar phenomenon has been reported for the case of graphene.41 As for pure MoS2, two broad peaks at 2θ = 32.8° and 57.3° can be indexed to the (100) and (110) diffraction planes of 2H-MoS2, while two peaks at 2θ = 9.6° and 18.6° indicate oxygen-incorporated MoS2 ultrathin nanosheets, which can dramatically enhance the HER activity compared with the pristine 2H-MoS2 (JCPDS Card No. 73-1508).42 As for the GNR@MoS2 hybrids, some characteristic diffraction peaks of MoS2 can be observed, while the absence of peaks at the low-angle region indicates that the MoS2 nanosheets are not restacked at all, which is advantageous for the HER activity.24 Besides, the absence of diffraction peaks for the GNRs in the hybrids may be ascribed to the relatively higher intensity of the diffraction peaks of MoS2.
To further confirm the reduction of the GNRs in the hybrids, Raman spectra are measured, as shown in Fig. S5.† The D band at 1363 cm−1 refers to a breathing mode of the κ-point photons of A1g symmetry while the G band at 1594 cm−1 corresponds to the first-order scattering of the E2g mode.43 It can be clearly observed that after reduction by N2H4, an increased D/G intensity ratio appears both in the curves of the pure GNRs and the GNR@MoS2-1 hybrid as compared with the GONRs, which is similar to what has been reported for graphene oxide.44 Thus, the increased number of small domains of aromaticity may be responsible for the enhanced D band.45
XPS spectra were further investigated to confirm the compositions and chemical states of the GNR@MoS2 hybrids. Fig. 5A reveals the co-existence of C, Mo, S and O elements in the products, without any detectable impurities. The high-resolution Mo 3d spectrum (Fig. 5B) presents two major peaks at 228.9 eV and 232.0 eV, corresponding to the binding energies of Mo 3d3/2 and Mo 3d5/2, characteristic of Mo4+ in MoS2, indicating the dominance of Mo(IV) in the GNR@MoS2 products.46 A small peak can be observed at 225.8 eV, which is attributed to the S 2s component in MoS2. Fig. 5C shows the high-resolution S 2p spectrum, and the peaks located at 162.9 eV and 161.8 eV refer to S 2p1/2 and S 2p3/2 orbits of divalent sulfide ions (S2−), respectively.47 The above results further prove the successful growth of MoS2 nanosheets onto the surface of the GNRs. Fig. 5D presents the high-resolution C 1s spectrum of the GNR@MoS2-1 hybrid, showing a standard carbon peak at 284.6 eV. The loading percentages of MoS2 in the hybrids can be estimated from the TGA curves (Fig. S6†), being 25.3%, 39.2% and 49.4% for the GNR@MoS2-0.5, GNR@MoS2-1 and GNR@MoS2-2 hybrids, respectively.
Fig. 5 XPS spectra of the GNR@MoS2-1 hybrid: (A) survey spectrum, and high-resolution (B) Mo 3d, (C) S 2p, and (D) C 1s spectra. |
Fig. 6 LSV polarization curves for the GNR@MoS2-0.5, GNR@MoS2-1 and GNR@MoS2-2 modified GCEs in N2-purged 0.5 M H2SO4 solution. |
The comparison of HER performance between the pure GNRs, MoS2, GNR&MoS2, MWCNT@MoS2 and the GNR@MoS2-1 hybrid is also investigated. As shown in Fig. 7A, the pure GNR sheets display an almost negligible HER activity with a horizontal line against the potential. Although both pure MoS2 and GNR&MoS2 exhibit distinct electrocatalytic ability, the current density and onset potential are far beyond those of the GNR@MoS2-1 hybrid, which may be due to the poor conductivity, limited exposed active edges and the insufficient contact between the GNR template and the electro-active MoS2 sheets. On the whole, the largely enhanced catalytic HER performance of the GNR@MoS2 hybrids can be attributed to the synergistic effect between the highly conductive GNRs and the electro-active MoS2 nanosheets, indicating the successful construction of the hybrid materials. In addition, it is worth mentioning that MWCNT@MoS2 exhibits significantly worse HER activity compared with that of GNR@MoS2-1, with a current density of 10 mA cm−2 at a voltage of −216 mV. The relatively poor catalytic performance for MWCNT@MoS2 may be due to the limited space of the outer wall of the MWCNTs for the growth of the MoS2 nanosheets, not taking full utilization of the inner surface or intra tube space, which can also be speculated from their morphological difference as discussed above. Moreover, MWCNTs with closed ends are also unfavorable for the electrolyte permeation, leading to inferior electrochemical activity.
Fig. 7 (A) LSV polarization curves for different material modified GCEs in N2-purged 0.5 M H2SO4 solution. (B) Tafel plots for the pure MoS2 and GNR@MoS2-1 modified GCE. |
The Tafel slope, an inherent property of the catalyst, is related to the reaction mechanism of a catalyst and determined by the rate-limiting step of the HER.11 The onset potential for the HER is determined from the semi-log plot, which is the starting point of the linear relationship of the Tafel curve. It is observed in Fig. 7B that the GNR@MoS2-1 hybrid displays superb HER activity with a low onset potential of −0.11 V, which surpasses that of pure MoS2. The Tafel slope can be deduced from the Tafel equation (η = blog(j) + a, where η is the overpotential, j is the current density, and b is the Tafel slope), yielding Tafel slopes of 43.4 and 82.3 mV per decade for GNR@MoS2-1 and pure MoS2, respectively.
In terms of the mechanism of hydrogen evolution, the Tafel slope of GNR@MoS2-1 indicates that the HER process takes place via a rapid Volmer reaction, followed by a rate-limiting Heyrovsky reaction.48,49 The HER performance of the GNR@MoS2 hybrids is comparable or superior to some previously reported MoS2-based HER catalysts, as listed in Table S1.† The extraordinary HER ability can be elucidated with the schematic diagram shown in Fig. 8. First, the GNR sheets not only serve as a conducting template, but also provide more surface area for the growth of MoS2, that is, both in the outer sheets and inner wall layers, as well as the edges along the quasi-one-dimensional strips. Second, the thin strips of quasi-one-dimensional GNRs are beneficial for the permeation of the electrolyte, which is rather different from the MWCNTs with relatively thicker walls. Third, the GNRs can form a highly conductive pathway, which facilitates the transportation of electrons. Last but not least, an appropriate amount of MoS2 nanosheets is decorated onto the GNRs through controllable adjustment of the morphology, thus effectively preventing the self-agglomeration of MoS2 and excessive exposure of the GNRs, taking full advantages of both components.
Electrochemical impedance spectroscopy (EIS) is a useful technique to evaluate the interface reactions and catalytic kinetics in the HER process. The Nyquist plots of the pure GNR, MoS2, MWCNT@MoS2 and GNR@MoS2-1 modified GCEs are displayed in Fig. 9A. The solution resistance Rs can be obtained from the intersection of the curves at the real axis in the range of the high frequency region while the charge transfer resistance Rct corresponds to the semicircle of the Nyquist plots. A straight line in the low frequency region and an inconspicuous arc in the high frequency region are shown in the Nyquist plot of GNR@MoS2-1, revealing a dramatically decreased Rs and Rct for the GNR@MoS2 hybrids compared with those for pure MoS2. It further implies that the GNRs can shorten the pathway of ion diffusion and facilitate the electron transfer, which gives prominence to the important function of a highly conductive template with a large surface area. For MWCNT@MoS2, it can be observed that its solution resistance Rs is apparently larger than that of GNR@MoS2-1, indicating poor contact between the electrolyte ions and hybrid material. More importantly, the semicircle of MWCNT@MoS2 is more obvious in comparison with that of the GNR@MoS2-1 hybrid (Fig. S7†), illustrating its higher charge resistance during the electrocatalytic process. As a result, the template of the GNRs indeed plays a critically significant role of facilitating the permeation of the electrolyte, which can be ascribed to the thinnish ribbon strips of the quasi-one-dimensional GNRs, providing an open space during the HER performance compared with the relatively closed structure of the muti-walled CNTs.
Fig. 9 (A) Nyquist plots for the GNRs, MoS2, MWCNT@MoS2 and GNR@MoS2-1 hybrid. (B) LSV polarization curves for the GNR@MoS2-1 modified GCE recorded before and after 1000 CV cycles. |
The durability of the GNR@MoS2-1 catalyst was also investigated using CV measurements by scanning 1000 cycles from −0.35 V to 0.25 V (vs. RHE) with a scan rate of 100 mV s−1 in 0.5 M H2SO4. As shown in Fig. 9B, it can be observed that the polarization curve after cycling displays a slight negative shift, which may be caused by the inevitable partial decrease of active edges of MoS2 for the HER.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra27180k |
This journal is © The Royal Society of Chemistry 2016 |