Andriana
Plantzopoulou
a,
Ioanna K.
Sideri‡
a,
Anastasios
Stergiou‡
a,
Martha
Kafetzi
a,
Stergios
Pispas
*a,
Raul
Arenal
*bcd and
Nikos
Tagmatarchis
*a
aTheoretical and Physical Chemistry Institute, National Hellenic Research Foundation, 48 Vassileos Constantinou Avenue, Athens 11635, Greece. E-mail: tagmatar@eie.gr
bLaboratorio de Microscopias Avanzadas (LMA), Universidad de Zaragoza, Mariano Esquillor s/n, 50018 Zaragoza, Spain
cInstituto de Nanociencia y Materiales de Aragon (INMA), CSIC-U. de Zaragoza, Calle Pedro Cerbuna 12, 50009 Zaragoza, Spain
dARAID Foundation, 50018 Zaragoza, Spain
First published on 2nd May 2022
Development of synthetic strategies towards two-dimensional laminar transition metal dichalcogenide/graphene heterostructures is of particular interest for the electrocatalytic hydrogen evolution reaction (HER). Herein, we describe the preparation of a molybdenum disulfide/graphene heterostructure, mediated by a bifunctional polystyrene-b-poly(acrylic acid) block copolymer. Surface engineering of graphene nanosheets by incorporation of ammonium-terminated organic chains facilitated attractive electrostatic interactions with the anionic polycarboxylate block of the copolymer, while the phenyl rings of the polystyrene block of the copolymer enabled the development of multiple van der Waals π–S interactions with the surface sulfur atoms of MoS2. The heterostructure exhibits excellent HER performance, as a result of the synergistic effect of 1T-MoS2 loading onto graphene and the latter's charge delocalization efficiency. An onset overpotential value of −0.18 V vs. RHE, a small charge transfer resistance of 36.2 Ω and a large electrocatalytic surface area of 205 cm2 were obtained, paving the way for the construction of advanced heterostructures based on graphene and transition metal dichalcogenides with improved electrocatalytic activity.
The superiority of 1T (metallic and octahedral) versus 2H (semiconducting and trigonal prismatic) polytype of MoS2 with respect to the HER is well-known,7,8 and arguably, the research interest is now primarily oriented to derivatives with enhanced 1T-MoS2 phases, holding the prominence as one of the best earth-abundant HER electrocatalysts. Although 1T-MoS2 shows intrinsic elevated electrocatalytic activity for the HER, owing to its exposed abundant sulfur edges at the periphery9 and to the scarce unsaturated Mo atoms of the basal plane,10 there is still room to further ameliorate its performance based on certain characteristics. Relatively lower conductivity, potential aggregation, thermodynamic instability of the 1T phase and the catalytically inert/scarcely activated basal plane of the material are some of these characteristics. To this end, efforts have been made to increase the active sites, their electrical contact, the mass and electron transfer and the specific surface area of the material by interfacing MoS2 with graphene-related materials. For example, growth of MoS2 nanopatches on single-walled carbon nanotubes was recently found to promote efficiently the electrocatalysis of the HER,11 as does a 2D/2D hetero-layered van der Waals architecture of MoS2/graphene.12 The stimulus, however, had been given earlier, with the growth of MoS2 nanoparticles on reduced graphene oxide (RGO), as a reliable solution to improve the electrocatalytic performance of MoS2via hybridization with graphitic nanosheets,6 and the promising HER results inspired other research groups to engage in this pairing.13–16 RGO was also utilized to tune the electronic states of MoS2 in MoS2/RGO, and the effect of the latter on the HER was evaluated.17 The HER electrocatalytic activity of MoS2 has reportedly been boosted by interfacing nitrogen-doped RGO nanosheets,18 while the coupling of MoS2 hollow spheres with RGO towards the HER electrocatalysis and energy conversion applications has been also presented.19 Recently, GO and RGO have been comparatively studied for the electrocatalysis of the HER in conjunction with MoS2, and the MoS2/GO heterostructure was found to be catalytically superior.20 In a recent study, the excellent HER performance of MoS2/GO was ascribed to the metallic 1T phase domains of MoS2-based materials and strong interactions between the two counterparts, albeit the mechanisms taking place remain quite unclear.21 However, all the aforementioned synthetic procedures involve the implementation of graphene derivatives and inorganic precursors of MoS2 by solvothermal/hydrothermal methods, high-pressure autoclave methods and exposure to microwave irradiation. Consequently and as a significant handicap, these methodologies do not provide control over the morphology and chemical composition of the hybrids, especially those of MoS2; therefore, there is limited insight not only regarding the chemical interaction of the counterparts, but also for the reasoning behind their synergism and thus the enhanced HER performance of the hybrids.
Yet, the different chemical structures of graphene oxide and exfoliated transition metal dichalcogenide nanosheets, that allow functionalization and introduction of specific functional groups, offer the tools for the design of heterostructures via specific and controlled covalent and non-covalent interactions. In this direction, we envisioned a rational approach for the construction of a novel layered MoS2/graphene heterostructure for the electrocatalysis of the HER, employing the bifunctional polystyrene-b-poly(acrylic acid), PS-b-PAA, block copolymer as an interfacial modifier, adhering the two laminar nanostructures and allowing charge transport. More specifically, we herein present a covalent functionalization route accessing positively charged graphene oxide nanosheets to enable attractive electrostatic interactions with the anionic polycarboxylate PAA block of the copolymer, while on the other hand, multiple van der Waals S–π interactions between surface sulfur atoms of chemically exfoliated MoS2 and the benzene rings of the PS block of the copolymer take place. This approach allows room-temperature processing of the MoS2/graphene heterostructure in the solution phase without self-aggregation of the counterpart nanosheets. The newly derived heterostructure was extensively characterized via IR and Raman spectroscopies, thermogravimetric analysis (TGA), ζ-potential and dynamic light scattering (DLS) measurements, (scanning) transmission electron microscopy ((S)TEM) imaging and energy dispersive X-ray spectroscopy (EDS), and the electrocatalytic HER performance and durability of the as-prepared heterostructure were also examined.
The PS macro-CTA is of high purity and free of any monomer, as witnessed by 1H NMR spectroscopy, and its molecular weight (Mw) was estimated through size exclusion chromatography to be 4.700 g mol−1 with a Mw/Mn = 1.19 (Fig. S1†).
The PS-b-PAA copolymer is of high purity, as witnessed by 1H NMR spectroscopy (Fig. S2†). The PS:
PAA ratio was also calculated via the 1H NMR spectrum via the H3–H4 protons of PS and H9 protons of PAA, as depicted in Fig. S2.† Calculations are explained in the ESI† section.
In a separate 10 mL flask, GO+ nanosheets (3 mg) and 5 mL DMSO were added and placed in a sonication bath for 1 h. This suspension was added to the ce-MoS2/PS-b-PAA mixture and was left under stirring for 3 days. The MoS2/PS-b-PAA/GO+ heterostructure was collected on a PTFE membrane filter (pore size = 500 nm) and washed with DMSO, distilled water and methanol to remove the excess copolymer. Finally, the heterostructure was dried in a vacuum chamber at r.t.
Initially, a simple 3-step synthetic route was employed to produce GO+ nanosheets (Fig. 1a). First, graphene oxide, conventionally produced from graphite, was treated with thionyl chloride to promote the conversion of the carboxylic acid functionalities to the corresponding highly reactive acyl chlorides. Then, the N-tert-butoxycarbonyl-2,2′-(ethylenedioxy)bis-(ethylamine) chain, abbreviated as EDBEA-BOC,22,23 was covalently attached via a condensation reaction, furnishing the GO-BOC derivative. Acidic hydrolysis of the BOC-protecting group afforded the target positively charged modified GO+ nanosheets (Fig. 1a).
The functionalization process was complementary monitored via Fourier transform infrared spectroscopy (FT-IR), thermogravimetric analysis (TGA) and spatial Raman spectroscopy. Oxidation of graphite was evident in the FT-IR spectrum of GO, where a dominant carbonyl vibration mode (1722 cm−1) of the generated carboxylic acid functionalities is observed (Fig. 1b). The incorporation of thermally unstable oxygen functionalities onto graphene was also probed during heating of the GO nanosheets under nitrogen, where a characteristic low-temperature thermal degradation occurs, in contrast to the thermally stable graphite starting material (Fig. 1c). Lattice defects across GO were also validated by the evolution of the D band in the Raman spectrum (Fig. 2a), being uniformly distributed in the sample, as revealed by spatial Raman mapping via calculating the D-to-G band intensity ratio, with an average value of 0.88 (Fig. 2b).
Noteworthily, heteroatom doping (oxygen functionalities) due to oxidation of graphene also affects the position of the G-band (27 cm−1 upshift). Having a uniform starting material, GO nanosheets, is essential to perform the designed chemical route towards the realization of positively charged GO+ analogues; otherwise, covalent functionalization will be partial or uncontrolled. The covalent incorporation of the organic chains onto GO was validated both in the stage of being BOC-protected, as well as in the final positively charged species at GO+. From the FT-IR spectrum of GO-BOC, we observe the carbonyl amide vibration mode at 1665 cm−1 and the N–H amide mode at 1519 cm−1, as well as the carbamide carbonyl mode of the BOC group at 1700 cm−1. Upon hydrolysis of GO-BOC towards GO+, the carbamide mode vanishes, while the amide-related modes remain intact, manifesting not only the completion of the reaction but also the presence of the covalently attached organic chains in the final material (Fig. 1b). Thermogravimetric analysis of GO-BOC and GO+ sheds more light on the chemical process, in terms of the chemical functionalization, as well as the impact of the chemical treatment on the nanosheets. Fast thermal degradation of EDBEA-BOC is observed at ∼200 °C, triggered by decomposition of the carbamide species (Fig. 1c). For both GO-BOC and GO+, the organic chains were degraded at higher temperature, gaining thermal stability due to the presence of graphene. It is noteworthy that, during the activation of the carboxylic acid functionalities using thionyl chloride, mild temperature annealing was used (see experimental procedures for details). This is depicted in the TGA graph, showing a slightly better stability of GO-BOC and GO+, and also in the Raman spectra. The D/G ratio for GO-BOC, as evidenced from the spatially collected spectra (Fig. 2c and d), is found increased (0.97), as compared to that of GO. Taking into consideration the employed amide-based chemistry, no further defects are expected in GO-BOC; thus, the D/G ratio should, in principle, remain constant. However, it is the use of the employed temperature in the chemical reaction, which initiated a healing process by partial degradation of the lattice, responsible for the witnessed increment of the D/G ratio in GO-BOC. Lattice healing is expected to increase the thermal stability of the nanosheets, which is evident for both GO-BOC and GO+, as compared to GO. This is further validated by comparing the results from TGA and the D/G ratio in the Raman spectra between GO-BOC and GO+. Hydrolysis of the BOC-protecting group is not expected to introduce defects in the graphene lattice, since the reaction proceeds at room temperature, and the chemistry occurs on the already covalently grafted organic chain. Actually, this was the case, and the D/G ratio in the Raman spectra is not altered as evidenced by the spatial Raman spectroscopy mapping of GO+ (Fig. 2d). Finally, the presence of free amines on GO+ was calculated via the ninhydrin (Kaiser) test24 (see the Experimental section) and found to be as high as 647 μmol of amines per gram of GO+ (Fig. S3†). Overall, the functionalization route towards GO+ is straightforward, based on cheap reagents and can be easily monitored, furnishing highly uniform modified nanosheets. In parallel, styrene was polymerized via reversible addition fragmentation chain transfer (RAFT) polymerization and furnished polystyrene macromolecular chains, with a molecular weight of 4.700 g mol−1 and a polydispersity of 1.19, as evaluated by size exclusion chromatography (SEC). The PS was then used as a macro-chain transfer agent (macro-CTA) for the growth of the polyacrylic block. The composition of the PS-b-PAA copolymer was found to be 33% PS and 67% PAA, with a molecular weight of 11.300 g mol−1 and high purity. Treatment of PS-b-PAA with an equimolar (with respect to carboxylate groups of the PAA block) amount of sodium hydroxide enabled the ionization of PAA, the realization of the negatively charged polyacrylate block and its electrostatic interaction with the previously synthesized positively charged GO+. Next, we focused on preparing ce-MoS2 of high quality and high 1T phase content. Our approach relies on a simple protocol based on conventional laboratory apparatus, not requiring the use of a dry box.25 A commercial semiconducting bulk MoS2 powder was treated with n-butyl lithium under nitrogen, followed by removal of the excess organolithium reagent and tip sonication of the organolithium-intercalated MoS2 material in distilled water. Accordingly, we managed to produce dispersions with a concentration of exfoliated nanosheets of up to 0.96 mg mL−1, which are stable for over a month if stored in the absence of light and oxygen. Symmetry transition from the semiconducting hexagonal prismatic 2H to the metallic octahedral 1T occurred as a result of the chemical treatment. The latter was probed by UV-vis spectroscopy, where the freshly prepared ce-MoS2 nanosheets lack the characteristic excitonic features of the semiconducting mode (Fig. S4†).26 Phase transition is further illustrated in the recorded Raman spectra at 514 and 633 nm excitation, where the characteristic J phonon modes of the 1T symmetry at 151, 224 and 330 cm−1 appear for ce-MoS2 (Fig. S5†). To promote the formation of the MoS2/graphene heterostructure, freshly-prepared ce-MoS2 aqueous suspensions were used in order to take advantage of the high stability of the nanosheets in aqueous media and the absence of aggregates. Tuning of the surface charges of GO+ is essential to construct the heterostructure with MoS2. Zeta (ζ) potential assays allowed us to estimate the charges present on the surface of graphene- and MoS2-derived nanosheets. GO displayed a ζ potential value of −29.7 mV, which is expected as a result of the carboxylic acid functionalities. Further, ce-MoS2 nanosheets also showed a negative ζ potential value of −29.1 mV, due to the chemical doping during the intercalation process. Evidently, these two materials are unlikely to form a heterostructure due to electrostatic repulsion. In contrast, the surface functionalization of GO towards the positively charged GO+ resulted in a positive ζ potential value of +17.3 mV, unlocking the formation of a heterostructure with MoS2 due to favorable electrostatic attraction. Herein, the role of the PS-b-PAA copolymer is to act as a bi-functional filler/linker, preventing the aggregation of the nanostructured counterparts and collapse due to the bulk nature of the nanosheets. In an alkaline environment, the ionized PAA block is capable to be immobilized on the surface of GO+via attractive electrostatic forces, while the PS block allows the development of S–π interactions between the sulphur atoms of MoS2 and the aryl-rings of the styrene functionalities. Accordingly, the GO+ and ce-MoS2 counterparts could form layered superstructures, namely MoS2/PS-b-PAA/GO+ (Fig. 3a), where the PS-b-PAA macromolecules could potentially occupy the interlayer space. Dynamic light scattering (DLS) studies provided information on the lateral size of GO+ and ce-MoS2 (Fig. 3b). In more detail, the hydrodynamic radius of the ce-MoS2 suspension was found to be 300 nm, with the corresponding value for GO+ being 700 nm. The formation of the MoS2/PS-b-PAA/GO+ heterostructure was investigated by mixing ce-MoS2 and GO+ counterparts (1:
1 mass ratio) in the presence of excess PS-b-PAA. Purification by vacuum membrane filtration was preferred, since removal of free PS-b-PAA and ce-MoS2 nanosheets was manageable via controlling the pore size of the membrane filter. In particular, the small lateral size ce-MoS2 nanosheets penetrate the membrane pores and are transferred to the filtrate, together with the excess PS-b-PAA, in contrast to the large lateral size GO+ remaining on the filter surface. Mixing of GO+, ce-MoS2 and PS-b-PAA in the wet phase resulted in a uniform stable dispersion, and upon filtration, the isolation of the heterostructure, stabilized by PS-b-PAA, was achieved. Formation of the MoS2/PS-b-PAA/GO+ heterostructure increased the hydrodynamic radius up to 1400 nm as a result of the produced laminated superstructure (Fig. 3b). The ζ-potential value of MoS2/PS-b-PAA/GO+ was significantly decreased (−17.0 mV) as compared to those of the two counterparts, indicating strong attraction between them within the isolated heterostructure. In contrast, in a reference material composed of pristine GO, ce-MoS2 and PS-b-PAA, we observed the formation
of aggregates during mixing, and the ζ-potential value of the isolated material remained highly negative (−24.7 mV), suggesting electrostatic repulsion. This was further elaborated by evaluating the reference material via Raman spectroscopy. Raman modes originating from ce-MoS2 were of low intensity, suggesting a non-uniform distribution of the nanosheets during mixing and isolation, implying the deficiency of GO to facilitate the formation of the heterostructure (Fig. S6†). The opposite is depicted in the Raman spectra (514 excitation) of the MoS2/PS-b-PAA/GO+ heterostructure (Fig. 3c), where more uniform distribution of the two materials within the heterostructure is evident, with both the Raman modes of ce-MoS2 and GO+ being intense. Noteworthily, the 1T phase of ce-MoS2 was found to be stabilized, since the characteristic J phonon modes are evident in all recorded spectra. This could be related to a delocalization of the electron surplus of ce-MoS2 (responsible for the 2H to 1T phase transition) over the PS block of the copolymer owing aryl units. Such a finding is of great importance with respect to the electrocatalytic properties of the heterostructure, since protection of the 1T phase is essential. In the FT-IR spectrum of MoS2/PS-b-PAA/GO+, the presence of PS-b-PAA in the anion (carboxylate) form is clearly observed (Fig. 3d).
Fig. 4a shows the HRTEM image of the MoS2/PS-b-PAA/GO+ heterostructure, showing the laminar nature of the structure with GO+ nanosheets being covered with MoS2 nanosheets. MoS2/PS-b-PAA/GO+ was further analyzed in detail by energy dispersive X-ray spectroscopy (EDS) in STEM. The heterostructure, composed of GO and MoS2 flakes, is clearly observed in the micrographs and EDS maps (Fig. 4b). Furthermore, two other areas of this MoS2/PS-b-PAA/GO+ flake have been also analyzed; one is mostly composed of GO+ (Fig. S7†), while in the second one, shown in Fig. S8,† MoS2 is the main phase. The low magnification STEM image of MoS2/PS-b-PAA/GO+ is also shown in Fig. S9.†
In the XPS survey spectrum of MoS2/PS-b-PAA/GO+, characteristic peaks due to S 2p, C 1s, Mo 3p, N 1s, and O 1s located at binding energies of ca. 164, 285, 395, 400 and 533 eV, respectively, are evident (Fig. S10†). The C 1s, O 1s, Mo 3d and S 2p core-level spectra are also shown in Fig. S10.† The contributions of Mo4+ 3d5/2 and Mo4+ 3d3/2 correspond, as expected, to the 1T-MoS2 phase.27 Regarding the analysis of the C 1s spectrum, the peaks at ∼284.5, ∼286 and ∼289 eV correspond to C–C, CO, and COOH, respectively, as it can be observed in GO samples.28
Next, the electrocatalytic behavior of the heterostructure was evaluated for the HER via LSV assessments in aqueous 0.5 M H2SO4. We screened the optimum mass ratio of the ce-MoS2 and GO+ counterparts within the heterostructure with that of ce-MoS2 and GO+ counterparts with the 10:
1 mass ratio being the best-performing and used for further evaluation of the electrocatalytic HER activity. Markedly, MoS2/PS-b-PAA/GO+ exhibited the lowest onset overpotential at −0.18 V vs. RHE, owing to 1T-MoS2 immobilized on the GO+ substrate, surpassing the electrocatalytic activity of both GO+ and ce-MoS2 reference materials (Fig. 5a). Furthermore, considering the functional current density required for sufficient hydrogen production at −10 mA cm−2, at which misinterpretations from inherent electroactivity can also be prevented, the potential registered for MoS2/PS-b-PAA/GO+ is −0.31 V, which is positively shifted by 130 mV compared to that of ce-MoS2 (i.e. −0.44 V) and by 370 mV compared to that of GO+ (i.e. −0.68 V). For comparison, the polarization curve of commercially available 20 wt% Pt on carbon black (Pt/C), the supreme catalyst for the HER, is also presented in Fig. 5a.
In the acidic-electrolyte-HER, the mechanism for electrocatalytic proton reduction and evolution of hydrogen is initiated by a primary discharge step (Volmer reaction) and is followed by either an electrochemical desorption (Heyrovsky reaction) or a recombination step (Tafel reaction). The discharge Volmer step involves the chemical incorporation of H into the catalyst by an oxonium ion and an electron association. During the Heyrovsky reaction, the H attached on the catalyst receives an electron from its surface and attacks a second oxonium ion to release gaseous H2 and H2O. Alternatively, H2 is produced by direct association of two H attached on the surface, during the recombination step.29 In order to determine which one of these steps is rate-limiting for the present electrocatalytic system, a Tafel post-experimental analysis was conducted. In Fig. 5b the Tafel slopes of MoS2/PS-b-PAA/GO+, ce-MoS2, GO+ and Pt/C, extracted from their respective LSV polarization curves, are depicted. The MoS2/PS-b-PAA/GO+ Tafel slope, which is 97 mV dec−1, is greater than that of Pt/C (i.e. 30 mV dec−1), but fairly lower than those of ce-MoS2 (i.e. 114 mV dec−1) and GO+ (i.e. 283 mV dec−1). Based on the calculated values, the HER of MoS2/PS-b-PAA/GO+ is governed by the Heyrovsky desorption.30 The easier kinetics of the HER for the heterostructure, reflected by the low Tafel slope, reveal the accessibility and therefore reactivity of protons on the electrochemically active sites.
The HER kinetics were further assessed by electrochemical impedance spectroscopy (EIS), by which charge-transfer phenomena at the electrode/electrolyte interface under HER conditions were evaluated. The EIS measurements were performed in the low overpotential (kinetic) region, and the obtained data were fitted to the Randles circuit.27 In more detail, the Nyquist plot of the MoS2/PS-b-PAA/GO+ heterostructure exhibited a charge transfer resistance (Rct) as small as 36 Ω, more than twice lower compared to those of ce-MoS2 (i.e. 69 Ω) and GO+ (i.e. 146 Ω) and only 5 times higher than that of Pt/C (Fig. 5c). The smaller the diameter of the semicircle of the Nyquist plot, that correlates the electron transfer from the electrode surface into the electrolyte, the less the charge transfer resistance is. Thus, the decrement of the diameter of MoS2/PS-b-PAA/GO+, by almost 50% as compared to that of ce-MoS2 and by 75% as compared to that of GO+, further justifies the beneficial effect of combining the two nanosheets, highlighting the superior electrocatalytic activity of the heterostructure. The latter was further evaluated in a complementary manner, from the electrochemical surface area values (ECSA), that were calculated based on the cyclic voltammograms at different scan rates. In particular, whilst GO+ is characterized by an extremely small (i.e. 13 cm2) and inefficient ECSA (Fig. S11†), MoS2/PS-b-PAA/GO+ exhibits a higher ECSA value of 205 cm2 (Fig. 5d), which is significantly greater than the ECSA of ce-MoS2 (i.e. 123 cm2) (Fig. S12†) and considerably closer to that of Pt/C (i.e. 362 cm2) (Fig. S13†). Taking into account the calculated ECSA values for each material, the ECSA normalized LSV polarization curves are depicted in Fig. S14,† highlighting the intrinsic electrocatalytic activity of the MoS2/PS-b-PAA/GO+ material.
To conclude the electrocatalytic studies, the long-term stability of the MoS2/PS-b-PAA/GO+ heterostructure was evaluated via both durability assessments and a chronoamperometry measurement. Fig. 5a depicts the LSV polarization curves of ce-MoS2, GO+ and MoS2/PS-b-PAA/GO+ after 10000 on-going catalytic cycles (dotted lines), where little overpotential increase is registered for the hybrid heterostructure (i.e. 12.5%). In order to highlight this important electrocatalytic aspect, we also recorded the chronoamperometric response of MoS2/PS-b-PAA/GO+ at a constant applied potential of −0.27 V vs. RHE in a time span of 10
000 s (Fig. S15a†). While the material was able to retain 65% of its original value after 10
000 s, we performed the complete electrocatalytic study afterwards to further validate this result. Interestingly, both the onset potential and the overpotential at −10 mA cm−2 shifted by only 10 mV (Fig. S15b†) (Table S1†). Likewise, the Tafel slope (Fig. S15b† inset) as low as 94 mV dec−1 was preserved. In contrast, the electrocatalytic activity loss upon recycling was evident from the EIS result, where the charge transfer resistance was found to be 38% higher (i.e. 59 Ω) than its initial (Fig. S15c†), but still lower than the respective ones of the reference materials (Table S1†). Finally, MoS2/PS-b-PAA/GO+ managed to retain more than 50% of its original ECSA after 10
000 s of use (Fig. S15d†).
Collectively, MoS2/PS-b-PAA/GO+ is an efficient electrocatalyst for the HER (Table S1†) owing to the synergistic effect of the two nanosheets/counterparts and to their unique chemical interactions. Graphene nanosheets provide high electrical conductivity and an increased surface area to not only facilitate the HER, but also to serve as a support matrix to MoS2.16 On the other hand, the high 1T-MoS2 content in the MoS2/PS-b-PAA/GO+ heterostructure ensures a plethora of active sites and the reduced charge-transfer resistance.7 In parallel, the chemical interaction between the two species prevents MoS2 aggregation phenomena and facilitates the electron flow by creating a vast electroactive surface area, while the PS-b-PAA bridge locks the electrocatalytically active 1T phase of MoS2.
Footnotes |
† Electronic supplementary information (ESI) available: Additional spectra, STEM images, voltammograms and HER parameters. See https://doi.org/10.1039/d2se00218c |
‡ Equally contributed to this work. |
This journal is © The Royal Society of Chemistry 2022 |