Introduction

The lack of a cost-effective replacement for Pt has plagued the scale-up of hydrogen electrochemical production \(\left( {2H^ + + 2{\mathrm{e}}^ - \to H_2} \right)\) for decades; the alternative catalytic materials are fundamentally limited by either a low catalytic efficiency or a short lifetime1, 2. Lamellar MoS2 has been regarded highly promising towards hydrogen evolution reaction (HER) since the activity of its metallic edges (ΔGH = 0.06 eV) was theoretically predicted by Norskøv and co-workers3, and was later on proved experimentally by Jaramillo and co-workers4, 5. The current guiding principles for advancing the MoS2 catalytic efficiency are as follows: First, increase the atomically undercoordinated active sites density in the trigonal prismatic phase (2H) MoS2, either through the preferentially exposing edge sites or through creating in-plane sulfur vacancies (SVs)6,7,8. However, unleashing the intrinsically high activity is retarded by the semiconductive feature of 2H–MoS2, where the charge transfer efficiency is limited by a deficiency of electrons at the reaction interface9, 10. Second, drive the 2H phase MoS2 into the conductive and therefore more catalytically active 1T phase11, 12. The basal-plane S atoms are regarded as active sites in 1T–MoS2; however, these S sites suffer from less favorable hydrogen adsorption features (ΔGH = 0.17 eV) despite the greatly increased site density13. Beyond the above mentioned problems of 2H–MoS2 and 1T–MoS2, one major issue that both these materials encounter is their reduced stability because defective 2H–MoS2 suffers from a high sulfur leaching rate14 and 1T–MoS2 is intrinsically metastable15. Apparently, MoS2 only become truly applicable towards the HER when the electronic conductivity, site density, intrinsic activity, and stability issues are simultaneously solved.

Here, we report a highly active and long-life MoS2-based HER catalyst, which is achieved by chemically activating its surface basal plane. We accomplish this by devising a thermodynamically spontaneous interfacial MoS2/Pd (II) redox reaction. Pd was atomically doped into the original Mo sites, causing the generation of SVs, the conversion to the stabilized 1T phase, the stabilization of defective sites, and the intrinsic activation of the 1T basal plane. The final Pd–MoS2 exhibits the highest HER performance ever achieved on heteroatom-doped MoS2-based materials in an acidic solution, along with good cycling stability and an exceptional anti-leaching feature that exceed those of undoped MoS2.

Results

Design of Pd–MoS2

We began by recognizing the redox characteristics of the chemically synthesized MoS2. Unlike the standard MoS2 samples, chemically synthesized MoS2 generally contains a certain concentration of defects, thus resulting a final Mo to S stoichiometry deviated from the theoretical ratio of 1:2. According to the principles of the conservation of charge in defect chemistry, a mixed valence can thus be created, thereby endowing MoS2 with redox power. We synthesized MoS2 through a typical wet chemical method (homemade, MoS2-HM, further denoted as MoS2 in the following, see the Methods section for the details). The sample was first characterized by inductively coupled plasma mass spectrometry (ICP-MS), which shows a Mo to S stoichiometry of 1:1.87 (Supplementary Table 1), corresponding to a Mo average valence state of 3.74. The X-ray absorption near-edge structure (XANES) test was further carried out, with the standard 2H–MoS2 used as a reference sample. The Mo L3-edge XANES results (Fig. 1a) demonstrate a decrease in the white line resonance strength in comparison to the standard 2H MoS2 sample, confirming a reduction in the unoccupied density of state (DOS) of Mo 4d and an average valance state lower than IV.16 The high-resolution Mo 3d X-ray photoelectron spectroscopy (XPS) measurement directly evidence the concurrent presence of Mo (III) and Mo (IV) (Fig. 1b) with deconvoluted doublets unambiguously assigned to Mo (III) (3d3/2 at 231.5 eV and 3d5/2 at 228 eV) and Mo (IV) (3d3/2 at 231.9 eV and 3d5/2 at 228.7 eV), respectively.

$$4{\rm{Mo}}^{3 + } \to 3{\rm{Mo}}^{4 + } + 3\mathrm{e}\left( {\rm{left}} \right) + V_{{\rm{Mo}}}'''({\rm{one}}\;{\rm{Mo}}\;{\rm{left}})$$
(1)
$$2{\rm{Mo}}^{4 + } + \mathrm{S}^{2 - } + 2\mathrm{e} \to 2{\rm{Mo}}^{3 + } + V_{\mathrm{S}}^{..}\left( {{\rm{one}}\;{\rm{S}}\;{\rm{left}}} \right)$$
(2)

The Mo (III)/Mo (IV) couple represents a standard electrode potential at –0.04 V. If properly paired with another redox couple, the redox reaction is expected to occur at the MoS2/liquid or gas interface. It is noted that structural vacancies can be created on the MoS2 surface during Mo redox process, according to the principles of defect chemistry, i.e., the conservation of mass and charge, and the host structure. As shown in Eqs. 1 and 2, while Mo (III) oxidation creates Mo vacancies, Mo (IV) reduction leads to the formation of SVs. Meanwhile, inter-valence charge transfer between two ions17 and abundant SVs18 can also induce the phase transformation into the 1T structure. After carefully screening the transition elements, we deliberately chose Pd (II) to accomplish the interfacial reaction with MoS2, as illustrated in Fig. 1c. A two-step thermodynamically spontaneous reaction is expected: First, the redox process in Eqn. 3 is a thermodynamically spontaneous reaction (\(E^\theta = 1.031\;{\mathrm{V}},\Delta rG^\theta = - 198.98\;{\rm{kJ}}\), Supplementary Note 1), which leads to the reduction of Pd and oxidation of Mo, therefore, creating Mo vacancies due to the principles of conservation of charge. Afterwards, however, metallic Pd is thermodynamically favorable towards anchoring to the energetic Mo vacancies and spontaneously forming the more stable Pd–S bond (Ksp=2.03 × 10−58, Supplementary Table 2), as shown in Eq. 4, through its incorporation into the MoS2 backbone (\(E^\theta = 0.75\;{\mathrm{V}},\Delta rG^\theta = - 144.75\;{\rm{kJ}}\), Supplementary Note 1). By injecting electrons into the MoS2 substrates, Mo (IV) is reduced back to Mo (III) and causes the leaching of S into the solution (law of conservation of charge) and the formation of S vacancies (see Eqs. 12). Thus, the MoS2 basal plane is driven to spontaneously incorporate Pd–S covalent bonds and form abundant SVs, presumably accompanied with phase conversion to form 1T-Pd–MoS2.

$$\begin{array}{l}2{\rm{Mo}}^{3 + } + {\rm{Pd}}^{2 + } \to 2{\rm{Mo}}^{4 + } + {\rm{Pd}}\\ \left( {E^\theta = 1.031\;{\mathrm{V}},\Delta rG^\theta = - 198.98\;{\rm{kJ}}} \right)\end{array}$$
(3)
$$\begin{array}{l}{\rm{Mo}}^{4 + } + {\mathrm{Pd}} + S^{2 - } \to {\rm{Mo}}^{3 + } + {\rm{Pd}}S\\ \left( {E^\theta = 0.75\;{\rm{V}},\Delta rG^\theta = - 144.75\;{\rm{kJ}}} \right)\end{array}$$
(4)
Fig. 1
figure 1

Design of a Pd–MoS2 material based on recognizing the redox characteristics of MoS2. a Mo L3-edge XANES spectra of the homemade MoS2. b High-resolution XPS results (Mo 3d region) of the homemade MoS2. c Schematic illustration of the spontaneous MoS2/Pd (II) redox reaction

Structure alteration due to Pd atomic do**

The redox power of MoS2 was first confirmed by a series of designed experiments between the pristine MoS2 sample and the Pd (II), Pt (IV), and Au (III) complex solutions, where the detailed results and explanations are shown in Supporting Information (Supplementary Note 2, Supplementary Figs. 1-2, and Supplementary Tables 3-4). For the final Pd–MoS2 samples, Pd with varied contents (1–15% Pd–MoS2) were expectedly immobilized as Pd(II) in MoS2, confirmed by the presence of the binding energy peaks at 336.7 eV (3d5/2) and 342 eV (3d3/2) (Fig. 2a). The Pd introduction does not induce observable morphological changes (Supplementary Figs. 3-4) to the MoS2 nanosheets, and is (Supplementary Fig. 5) found homogeneously distributed by high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) and elemental map**, with no Pd-based crystalline phases observed (Fig. 2b). The results from in situ heat-treatment TEM coupled with electron energy loss spectroscopy (EELS, Supplementary Fig. 6) and X-ray diffraction (XRD, Supplementary Fig. 7) tests suggest that Pd was firmly integrated into the MoS2 backbone without phase segregation even at 600 °C.

Fig. 2
figure 2

Structural characterization of 1%Pd–MoS2 and MoS2. a High-resolution XPS results (Pd 3d region) of the 1%Pd–MoS2. b XRD patterns. c Fourier transform of the k3-weighted Pd K-edge of the EXAFS spectra. d Fourier transform of the k3-weighted Mo K-edge of the EXAFS spectra. e Dark-field scanning transmission electron microscopy image of the 1%Pd–MoS2. Blue and yellow balls indicate Mo and S atoms, respectively. Scale bar: 1 nm. f ESR spectra. g Raman spectra of 1%Pd–MoS2 and MoS2. h Bode spectra obtained through electrochemical impedance spectroscopy with a frequency range from 0.1 Hz to 10 kHz and amplitude of 5.0 mV

To elucidate the Pd local bonding environment and the occupation sites in MoS2, extended X-ray absorption fine structure (EXAFS) and sub-angstrom resolution aberration-corrected HAADF-STEM were carried out. The Fourier transform of the k3-weighted Pd K-edge EXAFS spectra (Fig. 2c, R-space) for Pd–MoS2 shows the disappearance of the first-shell Pd–Pd scattering peak at 2.51 Å in comparison with Pd foils, indicating that the Pd species are formed neither as metallic Pd nanoparticles nor as Pd clusters. A prominent peak centered at a much lower R position is observed at 1.84 Å instead, corroborating the dominance of the Pd–S scattering contribution. We further fitted the main Fourier transform peaks from 1.1 to 2.5 Å in the R-space to quantitatively extract the Pd coordination parameters (Supplementary Fig. 8 and Supplementary Table 5). The best fit indicated a Pd–S bond distance of 2.33 Å and a Pd to S coordination number of 4.33, which is smaller than the nominal Mo–S coordination19. The Mo K-edge EXAFS spectra (Fig. 2d) reveal a decrease in the Mo–S and Mo–Mo peak intensities after Pd do**. While the former can be attributed to newly generated SVs, the latter may originate from the rearrangement of both atoms due to Pd fixation and the formation of SVs. The absence of first-shell Pd–Mo scattering in both Fig. 2c, d excludes the possibility that Pd was directly bound with Mo, thereby confirming that Pd does not occupy the S sites in MoS2. The sub-angstrom resolution aberration-corrected HAADF-STEM (Supplementary Fig. 9) shows an ordered MoS2 crystalline structure after Pd fixation, which is evident of the atomic dispersion of Pd. Thus, judging from the successful immobilization of Pd into the MoS2 frame, the only remaining possibility is that Pd replaces the Mo site, as we expected.

We next examined the influence of Pd atomic do** on the formation of S vacancies by comparing the 1%Pd-MoS2 with pristine MoS2. Electron spin resonance (ESR Fig. 2f) is utilized to detect the paramagnetic signals, and the ~337 mT (g = 2) signal intensity reflects the concentration of unsaturated sites with unpaired electrons. The pristine MoS2 demonstrates a relative intensity at 0.52 × 103 a.u. mg−1 sites, corresponding to the edges, crystallite interfaces, and in-plane SVs20. Remarkably, the 1%Pd–MoS2 exhibits a signal intensity (1.55 × 103 a.u. mg−1) approximately 3 times that of the MoS2, corroborating the formation of abundant SVs, in accordance with Eqn. 2 and the EXAFS results (with an in-plane Pd–S bond coordination number of 4.3 and a decrease in the scattering strength for the Mo–S path). We further employed positron annihilation lifetime spectroscopy (PALS, Supplementary Fig. 10 and Supplementary Table 6) to understand the form and content of the defects. τ1 is assigned to the lattice defects, such as structural vacancies (τv) or dislocation vacancies (τd), while τ2 is caused by clusters of vacancies. Notably, both τ1 and τ2 increased (Table 1) after 1% Pd do**, i.e., from 183.6 ± 5.3 and 355.5 ± 5.9 ps to 206.2 ± 4.7 and 384.6 ± 8.1 ps, respectively, which corresponded to an increase in the defect dimension. At the same time, the τ1 and τ2 relative intensities, denoted as I1 and I2, are also tuned, showing an increase in I1 from 49.1% to 57.5% and a decrease in I2 from 49.2% to 40.8%, which suggested that more single lattice SVs than vacancy clusters are introduced. ICP-OES results show that the final 1%Pd–MoS2 represents a stoichiometry of Pd0.02MoS1.82, with an S ratio significantly lower than that in the pristine MoS1.87 sample. According to the XPS surface analysis, the total amount of surface SVs created due to Pd do** is estimated to be 16.7% (Supplementary Table 7).

Table 1 Summary of the electrochemical and structural properties of 1%Pd–MoS2 and MoS2 catalysts

The impact of Pd on the phase conversion was further investigated. Careful examination of the (Fig. 2e, Supplementary Fig. 11 and Supplementary Note 3) reveals structurally distinct domains in the 1%Pd–MoS2 sample. A trigonal lattice structure corresponding to the 1T–MoS2 and the common honeycomb lattice area of the 2H–MoS221, 22 are both clearly visible, as indicated by the circles in Fig. 2e. However, the predominant fraction of 1T–MoS2 demonstrates the heterogeneous structure (2H + 1T) of the Pd–MoS2, and Raman spectroscopy provides direct evidence of the 2H to 1T phase conversion. Pristine MoS2 (Fig. 2g) exhibits two peaks at 378 cm−1 and 404 cm−1, attributable to the 2H phase vibrational configurations of the in-plane Mo–S phonon mode (E2g) and the out-of-plane Mo-S mode (A1g), respectively23,24,25. In contrast, 1%Pd–MoS2 exhibits new prominent peaks at 146, 278, and 332 cm−1, which were obviously associated with the 1T-MoS2 phonon modes, along with distinctly weakened 2H–MoS2 signals (E2g and A1g). As a result of the phase conversion, the electronic conductivity (Fig. 2h, see the Supplementary Methods for details) is more than one order of magnitude greater for 1%Pd–MoS2 (2.619 S cm−1 versus 0.178 S cm−1 for MoS2) and approached that of 1T–MoS2 (10–100 S cm−1).26

The HER catalytic behavior

Next, we examined the HER catalytic behavior of Pd–MoS2 (1–15% do**), MoS2, Pd/C, and commercial Pt/C catalysts with the representative linear sweep voltammograms (LSVs) summarized in Fig. 3a and Supplementary Fig. 12a-b. First, the pristine MoS2 exhibited an overpotential of 10 mA cm−2 (η@10 mA cm−2) at 328 mV, being consistent with those reported for 2H–MoS2 in the literature27, 28. Second, Pd do** leads to a breakthrough in the catalytic performance towards the HER, far exceeds that of the metallic Pd catalysts. The 1%Pd–MoS2 exhibited a current density of 10 mA cm−2 at an overpotential of only 89 mV. This result corresponds to the highest performance ever reported for heteroatom-doped MoS2-based catalysts in acidic media28,29,30,

Fig. 3
figure 3

Superior activity and stability of 1%Pd–MoS2. a LSV polarization curves of MoS2, 1%Pd–MoS2, 1%Pd–MoS2/CP, 1%Pd–C, and 20%Pt–C (with iR correction). b Tafel plots derived from the results given in Fig. 3a. c Exchange current density for the MoS2, 1%Pd–MoS2, 1%Pd–C and 20%Pt–C samples, derived from the Tafel plots, as indicated by the dashed lines. d EIS comparison for the MoS2 and 1%Pd–MoS2 in terms of Nyquist plots; inset, the full-spectrum for the MoS2. e Stability measurements for MoS2 and 1%Pd–MoS2 using accelerated degradation tests (5000 cycles, 100 mV s−1); polarization curves are shown without iR correction. f ICP-OES results of dissolved S, Mo, and Pd ions in the electrolyte after the stability tests shown in Fig. 3e

The corresponding Tafel plots (Fig. 3b and Supplementary Fig. 12c) show that Pd atomic do** decreases the Tafel slope from 157 to 62–80 mV dec−1 (Pd–MoS2, 1–10%), demonstrating the transition of the rate determining step (RDS) away from the Volmer discharge reaction (H3O+ + e → Hads + H2O). This transition is expected since Pd do** results in a phase conversion to 1T–Pd–MoS2, thus making access to the electrons and the formation of Hads easier at the interface. The exchange current densities (j0) were further calculated to demonstrate the inherent HER activity, as shown in Fig. 3c. Pd do** is extremely effective for boosting the j0 because 1%Pd–MoS2 (805 µA cm−2, the best of all reported MoS2 based materials) is significantly better than MoS2 (37.25  µA  cm−2) and almost reached that of 20% Pt/C (918  µA  cm−2) at the same catalyst loading7, 27, 33. The electrochemical impedance spectroscopy (EIS) results explain the exceptional HER behavior of 1%Pd–MoS2, where the Nyquist plots (Fig. 3d) show a large reduction in the charge transfer resistance (Rct) from 92.89 Ω cm2 for MoS2 to 1.50 Ω cm2 (Supplementary Fig. 13 and Supplementary Table 9) for 1%Pd–MoS2.

In addition to having a high HER activity, 1%Pd–MoS2 is a stable and long-life catalyst. We combined the electrochemical measurements and ICP-OES test to verify the material durability during operation. The chronoamperometry test results (Supplementary Fig. 14) manifest that the 1%Pd–MoS2 exhibits an outstanding long-term operational stability beyond 100 h with an observed potential increase of only 14 mV. The ultrahigh stability of 1%Pd–MoS2 was also proved by long-term cyclic voltammetry tests, where no obvious potential decay was observed (Fig. 3e) after 5000 cycles. In contrast, the MoS2 electrode shows that η@10 mA cm−2 increased by 10 mV after 5000 cycle tests. The ex situ ICP-OES results present a reduced sulfur concentration (2.724 versus 2.291 ppm) in the testing electrolyte (Fig. 3f), which indicated that the final defective 1%Pd–MoS2 surface is even more stable than that of the pristine MoS2. Moreover, we further examined the Pd–MoS2 catalyst using XPS characterization after the above electrolysis test (Supplementary Fig. 15). Neither the content nor the state of Pd was altered, suggesting that Pd is firmly integrated into the MoS2 backbone and highly stable under electrolytic conditions. This result is contradicted with those reported in the literature because highly active and defective catalysts are always accompanied by poor stability. Thus, by tailoring the chemical bond characteristics, we designed a MoS2 material which simultaneously possesses good stability and activity.

Density functional theory calculations

Density functional theory (DFT) calculations were carried out to obtain atomic-scale insight into the do** effect of Pd. First, the energies of the Pd atoms on varied sites were calculated, and it is found that Pd exhibits a strong tendency to replace Mo with an exothermic energy of −2.22 eV compared to replacing S (−0.17 eV) adsorbed on the Mo atop site (1.75 eV) and the hollow site (2.35 eV) (see Supplementary Note 4 and Supplementary Fig. 16), thus supporting our expectation of the thermodynamically driven formation of the Pd–Mo–S x compound. Second, we calculated the energy for the formation of SVs in MoS2 and Pd–MoS2 (Fig. 4b and Supplementary Fig. 17), and the energy for SVs formation decreased by ~1–2 eV due to the Pd do**. Thus, we can use the spontaneous Pd do** strategy to create a large number of SVs on the MoS2. Third, we explored the influence of Pd do** and the SV concentration on the total phase energy of both the 1T and 2H MoS2. In Fig. 4a-b, we found that 1T–MoS2 becomes more stable than 2H–MoS2 (see Supplementary Note 5 for the detailed results) with the presence of the Pd and SVs at certain concentrations. Specifically, the 1%Pd-MoS2sample, as confirmed by the XPS (Supplementary Table 7) to possess surface concentrations of Pd at 3.47% and newly generated SVs at 16.7%, corresponding to the higher stability of the 1T versus 2H phase shown in Fig. 4b. Finally, we investigated the effect of Pd do** on the HER activity of the MoS2 basal plane. The hydrogen adsorption free energy (ΔGH) was used to evaluate the hydrogen evolution activity. The Pd sites themselves were calculated to be inactive as H does not form a very stable adsorption structure on Pd atop site (Supplementary Fig. 18). The ∆GH of the SVs in the 1T MoS2 is 0.09 eV, suggesting more favorable HER catalytic behavior than that of 1T basal-plane (∆GH = 0.17 eV). Therefore, the increases in SVs concentration unambiguously contribute to the increased catalytic behavior, and we regard this as an increase in the site density. More excitingly, the ΔGH of the S atop site adjacent to Pd (Pd–S*–Mo) in the 1T–Pd–MoS2 exhibits an almost thermoneutral value of −0.02 eV, which was far better than the reported most favorable edge sites (Fig. 4c, d and Supplementary Fig. 19)4, 34. In Fig. 4e, the difference in the S atom electronic structures of site 1 (Pd adjacent S site) and site 2 (regular in-plane S site) were elucidated by the projected p-orbital DOSs of S. Because the ionization energy of Mo is less than that of Pd, site 2 receives more electrons from the adjacent three Mo atoms with more p-states filled, which leads to a weaker interaction with hydrogen35,Full size image