The tubular MoP/MoS2 composite assembled from nanosheets was synthesized by partially phosphating of tubular MoS2. Figure 1a shows the XRD patterns of MoP/MoS2-1. It is obvious that the diffraction peaks at 14.1o, 33.0 o, 39.5 o and 58.6 o are corresponding to (002), (100), (103) and (110) planes of typical MoS2 (JCPDS card no. 65-1951), respectively. While the diffraction peaks at 27.8 o, 32.0 o, 42.9 o, and 57.0 o can be ascribed to the (001), (100), (101), and (110) planes of MoP (JCPDS card no. 89-5110), respectively. So it is confirmed that the MoP can be successfully produced by the replacement of S atoms with P in MoS2, because they have similar atomic radii. SEM images of MoP/MoS2-1 are shown in Fig. 1b and c. The hollow tubes are approximately 10 µm in length and 1 µm in width, and the thickness of the wall is approximately 200 nm, which can be observed from the open end of a single tube. The walls of the hollow tubes are fabricated by loosely stacked nanosheets with the thickness of the nanosheet about 8 nm. It is obvious that the basal planes and the edges of the MoP/MoS2 composite can be totally exposed to the electrolyte. Compared with the MoS2 template as shown in Fig. S1a, the tubular structure and the nanosheets were well kept without any damage. The energy-dispersive X-ray (EDX) elemental mapping images of MoP/MoS2-1 are displayed in Fig. 1d. It can be observed that the Mo, S, and P elements are distributed homogeneously over the tubular MoP/MoS2-1. The hollow tubular structure of MoP/MoS2-1 assembled from nanosheets can be further confirmed by TEM images of MoP/MoS2-1 with different magnifications in Fig. 1e and f. A distinct heterojunction can be found in the HRTEM image as shown in Fig. 1g. The lattice spacing of 0.63 nm and 0.28 nm can be ascribed to the (002) plane of MoS2 and the (100) plane of MoP, respectively. The hollow tubular structure of MoP/MoS2-1 is favorable for the abundant mass diffusion, the nanosheets can provide active sites and short diffusion path, and the rich heterojunctions between MoS2 and MoP on the nanosheets may be favorable in promoting the electrocatalytic performance for HER because the migration of electrons become easily [30,31].
X-ray photoelectron spectroscopy (XPS) was conducted to analyze the chemical compositions and states of the MoP/MoS2-1. Figure 2a shows the survey XPS spectrum, and it is clear that MoP/MoS2-1 comprises Mo, S, and P elements. The XPS spectrum of Mo 3d can be divided into four peaks as shown in Fig. 2b. The peaks at 232.1 and 229.0 eV can be ascribed to the bonding energy of Mo 3d3/2 and Mo 3d5/2 of the Mo4+ [36], respectively. While the peaks at 231.1 and 227.8 eV correspond to the bonding energy of Mo 3d3/2 and Mo 3d5/2 of Mo3+. Figure 2c shows two distincted peaks at 162.9 and 161.7 eV, which are correspond to S 2p1/2 and 2p3/2 of MoP/MoS2-1, respectively. Figure 2d shows the P 2p spectra of MoP/MoS2-1 with two characteristic peaks at 130.0 eV for P 2p1/2 and 129.1 eV for P 2p3/2 of P-Mo bonds, respectively. These results further confirm that the coexistence of MoP and MoS2 in hierarchical tubular MoP/MoS2-1.
In order to get the optimal electrocatalyst for HER, a series of composites with different ratios of MoP to MoS2 were prepared by adjusting the weight ratio of NaH2PO2 to MoS2. Figure 3 shows the XRD patterns of MoP/MoS2-0, MoP/MoS2-1/3, MoP/MoS2-5 and MoP/MoS2-40. It is can be observed that the intensities of the diffraction peaks of MoP become stronger with the increase of the weight ratio of NaH2PO2 to MoS2. Especially when the 40 times of NaH2PO2 to MoS2 was used in the synthesis process, the diffraction peaks of MoS2 totally disappeared, and only those of MoP can be observed. This result indicates that the ratio of MoP to MoS2 can be adjusted by simply changing the weight ratio of NaH2PO2 to MoS2. The phosphorization degree can be reflected by the intensity ratios of the (100) diffraction plane of the MoP and MoS2 phase [30], and they are about 0.6, 1.6 and 2.1 for MoP/MoS2-1/3, MoP/MoS2-1 and MoP/MoS2-5, respectively.
The SEM images of the samples were shown in Fig. 4. It is obvious that MoP/MoS2-1/3 and MoP/MoS2-5 had almost the same tubular structure as MoS2 with the wall assembled from nanosheets, while the MoP/MoS2-40 was hollow tubes with the wall composed of nanoparticles. Figure S2 and S3 show the homogeneous distribution of Mo, S and P elements over MoP/MoS2-1/3 and MoP/MoS2-5. While only Mo and P elements can be observed on MoP/MoS2-40 (Fig. S4). The EDX results showed that the phosphorus content of the synthesized materials increased with the increase of the mass ratio of NaH2PO2 to MoS2 (Table S1). Figure 3a shows the survey XPS results of the MoP/MoS2-0, MoP/MoS2-1/3, MoP/MoS2-5 and MoP/MoS2-40. For MoP/MoS2-0, two peaks at 229.7 and 232.8 eV indicate the chemical state of Mo is +4, and two characteristic peaks at 162.6 eV and 163.6 eV correspond to S. MoP/MoS2-1/3 and MoP/MoS2-5 showed the similar spectrum with MoP/MoS2-1 except for the increase P/S atom ratio as shown in Fig. 3, and the calculated P/S atom ratio are 0.059, 0.088 and 0.114 for MoP/MoS2-1/3, MoP/MoS2-1 and MoP/MoS2-5, respectively. For MoP/MoS2-40, the peaks at 231.1 and 227.8 eV correspond to the bonding energy of Mo 3d3/2 and Mo 3d5/2 of Mo3+, respectively. The peaks at 130.0 eV for P 2p1/2 and 129.1 eV for P 2p3/2 of P-Mo bonds, respectively. According to the ratio of Mo3+ and Mo4+ in the composites, the ratio of MoP/MoS2 are calculated to be 0.041, 0.082 and 0.128 for MoP/MoS2-1/3, MoP/MoS2-1 and MoP/MoS2-5, respectively. Both XRD and XPS results confirmed that the increase of the MoP content with the increase of the weight ratio of NaH2PO2 to MoS2. The nitrogen sorption isotherm and the pore size distribution curves of the samples are shown in Fig. S5. The BET specific surface area of MoS2, MoP/MoS2-1/3, MoP/MoS2-1, MoP/MoS2-5 and MoP/MoS2-40 are 9.5, 13.1, 14.3, 14.3 and 7.1 m2/g. The higher specific surface area of MoP/MoS2-1 and MoP/MoS2-5 will provide more active sites for electrochemical reaction.
The HER activities of MoS2, MoP and a series of MoP/MoS2 composites were evaluated in a three-electrode electrochemical cell. As the control, the electrocatalytic performance of Pt/C was tested. Figure 5a shows the polarization curves of the samples in N2-saturated 0.5 M H2SO4. It is obvious that Pt/C displays superior HER electrocatalytic performance with a nearly zero onset potential and a high current density. Compared to other samples, MoS2 displayed the poorest activity with onset potential of 165 mV (obtained at 1 mA/cm2) and an overpotential of 260 mV to obtain a current density of 10 mA/cm2. The MoP/MoS2 composites and MoP displayed better electrochemical performances compared with the tubular MoS2 as shown in Fig. 5a. The onset potentials are about 71, 37, 51, and 56 mV for MoP/MoS2-1/3, MoP/MoS2-1 and MoP/MoS2-5 and MoP/MoS2-40, respectively. Among the catalysts, the activity MoP/MoS2-1 exhibited the highest activity for HER with an overpotential of 101 mV to deliver a current density of 10 mA/cm2. Though MoP/MoS2-1/3 and MoP/MoS2-5 are composed of MoP and MoS2, their performance for HER were poorer than MoP/MoS2-1 as shown in Fig. 5a, which may be ascribed the proper ratio of MoP and MoS2 in MoP/MoS2-1, and similar results have been reported [27,30,31].
Tafel slopes are commonly used to illustrate the inherent properties of the catalyst and reaction kinetic mechanism for HER, which are fitted to Tafel equation η=b log j+a (b is the slope and j is the current density). Figure 5b shows the Tafel slopes of Pt/C and the as-prepared samples, The Tafel slopes of Pt/C was 32 mV/dec, in accordance with reported values [30]. The Tafel slopes of MoP/MoS2-0, MoP/MoS2-1/3, MoP/MoS2-1, MoP/MoS2-5 and MoP/MoS2-40 were 84, 64, 56, 64 and 73 mV/dec, respectively. The slope values of the samples are in the range of 50–90 mV/dec which suggests that the HER taking place on the surfaces of the catalysts. MoS2, MoP/MoS2 composites and MoP should undergo the Volmer−Heyrovsky mechanism and the electrochemical desorption is the rate-determining step.
In order to study the intrinsic performance of the synthesized samples, the exchange current density (j0) was obtained by extrapolation from the Tafel plots. As shown in Fig. 5c, the j0 values are 0.011, 0.083, 0.247, 0.132 and 0.232 mA for MoS2, MoP/MoS2-1/3, MoP/MoS2-1, MoP/MoS2-5 and MoP/MoS2-40, respectively. It is obvious that MoP/MoS2-1 exhibited a higher j0 than the other samples, which demonstrates the fastest reaction rate per surface area on MoP/MoS2-1 electrode. In order to assess the electrochemical activity surface area (ECSA), the double layer capacitance (Cdl) at the catalyst–electrolyte interface was calculated using cyclic voltammetry plots in a potential range (0.1–0.2 V vs. RHE, in which no faradic reaction was observed) with various scan rates (10 to 60 mV/s). The typical CV curves of the MoP/MoS2-1 electrode are shown in Fig. 5d. The calculated double layer capacitance values of the MoP/MoS2-0, MoP/MoS2-1/3, MoP/MoS2-1, MoP/MoS2-5 and MoP/MoS2-40 were 2.8, 5.7, 7.6, 6.0 and 3.8 mF/cm2, respectively (Fig. 5e). The Cdl value of the MoP/MoS2-1 electrode is larger than those of the other samples, which indicates MoP/MoS2-1 electrode has a large ECSA with rich active sites towards the HER. As a HER catalyst, MoP/MoS2-1 outperformed the other as-prepared samples and recently reported molybdenum-based and non-noble-metal catalysts due to its low overpotential and small Tafel slope [18,37-39]. The long-term durability of MoP/MoS2-1 is evaluated by the long-term CV cycling and the time dependence of the current density at an overpotential of 100 mV. As shown in inset of Fig. 5f, the polarization curves of MoP/MoS2-1 after 500 CV cycles almost overlaps with that before the cycle. Furthermore, the current density slightly declined during the chronoamperometry operation for 12 h in 0.5 M H2SO4 (Fig. 5f). All the above results indicate that the catalyst has good cycle stability.
Generally, the HER activity of the catalyst is closely related to the relative free energies of H* absorption on the catalyst (ΔGH*). So density functional theory (DFT) calculations were conducted to get the adsorption energies of H* on the surfaces of MoS2, MoP/MoS2 composite and MoP to understand the activity origin of the catalyst. According to Sabatiers’ principle, the closer to zero the ΔGH* is, the higher HER activity it will have [40]. Figure 6 shows the geometric models of H* adsorbed on catalysts. The calculated ΔGH* for H* adsorption on MoS2 and MoP are 2.20 eV and −1.15 eV, respectively. While the value for absorption on MoP/MoS2 composite is about −0.35 eV. It can be concluded that the more favorable H* adsorption kinetics on the surface of the MoP/MoS2 compoaite during the HER process.
A good catalyst should have excellent HER activity in a wide pH range. To explore the application of the hierarchical tubular MoP/MoS2-1 composite under alkaline and neutral conditions, the polarization curves and Tafel plots of the samples were tested in 1 M KOH (Fig. S7 a,b) and 1 M PBS (Fig. S7 c,d). The Tafel slopes of MoP/MoS2-1 under alkaline and neutral conditions were 78 and 93 mV/dec, respectively. The current densities of the catalyst can reach 10 mA/cm2 at overpotential of 210 mV in 1 M KOH and 257 mV in 1 M PBS, respectively. The results demonstrate that hierarchical tubular MoP/MoS2 composites are promising catalysts for the HER over a broad pH range.