W 2 TiC 2 MXene and Co/W2TiC2 catalysts. The synthesis procedure of W2TiC2 MXene is shown in Fig. 1. The bulk W2TiAlC2 MAX precursor was first prepared by spark plasma sintering of commercial powders. Two dimensional W2TiC2 was then obtained by selective etching of the aluminum layers with hydrofluoric acid (HF). Tetramethylammonium hydroxide (TMAOH) was used followed by sonication to exfoliate the MXene to thinner layers. The X-ray diffraction (XRD) pattern of W2TiAlC2 showed the characteristic peak of (004) plane at 2θ = 19° (Fig. 1b), and the peak positions exhibited good agreement with the pattern simulated in VESTA 52 by importing the lattice parameters derived by the 2θ values of the (110), (101), and (004) peaks of the experimental results. XRD refinement was conducted by GSAS 53 software and the alignments of main peaks including (006), (103), (104), (008) and (106) were achieved, confirming the expected W2TiAlC2 composition and structure. The (002) diffraction peak showed up in a lower angle (2θ = 6.8°) after the HF etching, indicating a larger c lattice parameter of W2TiC2 due to the removal of aluminum layers. After exfoliation, the (002) plane peak shifted further to 2θ = 5.85°, consistent with the expansion along the [001] direction. The experimental XRD also aligned with VESTA simulated XRD for W2TiC2 (Supplementary Fig. 1d), showing the c lattice parameter increasing from to 1.873 nm to 3.009 nm. Scanning electron microscopy (SEM) image of W2TiC2 MXene showed a morphology of stacked nanosheets (Fig. 1d). More magnified observation through transmission electron microscopy (TEM) revealed that the nanosheets were composed of few layers of MXenes, as shown in the side view (Fig. 1e) by the contrast of light and shaded regions changing with number of MXene flakes. Lateral structures of alternating atom layers of various contrast were also observed from the sidelined pieces, further validating the M3C2 structure of W2TiC2.
Co was then loaded on the W2TiC2 MXenes by incipient wetness impregnation of cobalt chloride precursor, followed by reduction under H2/N2 atmosphere at different temperatures, denoted as Co/W2TiC2-n (n is applied temperature in degree Celsius, n = 500, 600, 700). The X-ray diffractograms of the Co/W2TiC2-500, 600 and 700 catalysts (Fig. 2a) exhibited broad small Co peaks at around 2θ = 44° region, suggesting the existence of metallic Co. The broadening of the peaks were resulted from small nanoparticle sizes, which was confirmed later by the Extended X-ray absorption fine structure spectroscopy (EXAFS) fitting results. The decrease in the intensity of the MXene peaks and slight shift to lower 2θ of the (002) peak with increasing temperature suggested a slight decrease in the inter-flake spacing after the introduction of Co. The SEM image in Fig. 2b showed that the thin layer morphology of the MXene was maintained after the Co loading with no large Co grain being observed, indicating the uniform distribution of Co on the MXene surface. The TEM image of Co/W2TiC2-500 showed the MXene layers, with a consistent d spacing measured as 0.24 nm for the (01-10) lattice planes. The selected area electron diffraction (SAED) pattern further confirmed the basal plane hexagonal symmetry structure (Figs. 2c, d). Owing to the high extra-nuclear electron density of tungsten element, Co imaging was hindered by W2TiC2 under TEM and cannot be observed clearly except by mass thickness contrast on very thin areas of sample edge. In Fig. 2c, small dark patches were found at the edge of a MXene flake. Similar dark regions of 1-2 nm size can be observed on very thin regions of the Co/W2TiC2-500 sample (Supplementary Fig. 3). Lattice fringes distinguished from the lattice of W2TiC2 were marked and measured to be 0.204 nm and 0.177 nm, matching the (111) and (200) planes of face-centered-cubic Co respectively. The grain boundaries were not sufficiently distinct owing to the same reason mentioned above, necessitating the employment of additional characterization techniques to validate the presence of Co particles or patches. The aforementioned small dark regions were not observed in Co/W2TiC2-600 and Co/W2TiC2-700 samples, which could potentially be attributed to the further reduced size of Co in these catalysts. High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) with energy dispersive X-ray spectroscopy (EDS) was employed to characterize the Co/W2TiC2 catalysts (Figs. 2e, f). The EDS mappings show the signals of W, Ti and Co uniformly distributed through the sample without obvious phase segregation. The two-dimensional structure of W2TiC2 featuring orderly arranged metal layers was clearly presented from the side view STEM image (Fig. 2e). As the contrast in STEM images was directly related to the atomic number of the elements, the outermost layers with brighter contrast were identified as W, while the inner layers with slightly darker contrast were attributed as Ti. The ordered three metal layers agreed well with the simulated W2TiC2 structure.
The structure of Co/W2TiC2 and the impact of reduction temperature on the metal-support interaction were further investigated by X-ray photoelectron spectroscopy (XPS) and ion scattering spectroscopy (ISS). The ex-situ XPS survey and high-resolution spectra were shown in Supplementary Figs. 4 and 3a-c. The survey spectra confirmed the presence of W, Ti, C, O, and Co. The W 4f spectrum displayed characteristic peaks at around 32.3 eV, corresponding to W in MXene environment (W-C). The slight shift to higher binding energy compared with pure tungsten is attributed to the redistribution of electron density between tungsten and carbon, resulting in the electron density of tungsten to decrease. Conversely to this trend, the binding energy of W-C decreases from 32.3 eV of W2TiC2 to 32.2 eV, 32.1 eV, and 32.1 eV for Co/W2TiC2-500, 600, and 700 samples, respectively, indicating more electrons drawn from Co towards W, which suggested an enhanced interaction between Co and surface W as the reduction temperature increased. The Co 2p spectrum showed characteristic peaks of metallic Co, Co-Ox and a pair of shake-up satellite of Co-Ox 2p3/2 and 2p1/2, indicating the presence of metallic Co and slightly oxidized surface. The Co peaks shifted slightly to higher binding energy from 777.5 eV to 777.8 eV and 778.0 eV (Co/W2TiC2-500, 600, and 700, respectively), which is consistent with the trend observed in the W spectrum, agreeing with the more electrons drawn from Co with increased reduction temperature and stronger metal-support interaction. Same positive binding energy shift of Ti with the increasing temperature further validated the regulation of reduction condition on the electronic structure and interaction.
Ion scattering spectroscopy (Fig. 3d) revealed the continuous scattered ions intensity profile change on the Co/W2TiC2-700 catalyst with the helium (He+) exposure time. At t = 0s, only Co and W peaks were observed with a barely noticeable Ti peak, indicateing the position of Co and W on the top surface of Co/W2TiC2. With the increasing exposure time, the He+ beam gradually etched the surface and the Ti peak became stronger, while the peak of Co gradually vanished. As ISS provided elemental information from the top monolayer of the surface, this depth profile further confirmed the presence of Co on the surface as well as the Ti sandwiched between W atom layers in the W2TiC2 MXene.
Co structure transformation between single sites and nanoparticles on W 2 TiC 2 . XANES and EXAFS studies were conducted to gain more accurate information about the structures and coordination environments of Co under different reduction temperatures. The Co K-edge XANES energy was the electromagnetic transition from 1s electron to the 4p empty orbital. The XANES energy is determined fromthe first inflection point of the leading edge. For 3d transition metals, a pre-edge peak is often observed, which is due to the 1s to 3d forbidden transition. The oxidation state could be determined by pre-edge peak and edge energy 54. Co foil (Co0) has no pre-edge peak and XANES energy is 7.7090 keV, while CoO reference (Co2+) has a pre-edge peak at 7.7084 keV and XANES energy at 7.7166 keV (Supplementary Fig. 5a). The edge energy of all our Co/W2TiC2 samples were very similar and nearly identical to that of Co0, indicating the metallic nature of Co (Supplementary Table 2). The similarity in the shapes of the XANES spectra indicated that the local structures on the catalysts are also similar (Fig. 3e).
Fitting the EXAFS spectra gave information about the coordination numbers (CNs) and bond distance, as summarized in Supplementary Table 2. The Co K-edge EXAFS spectra for Co foil and CoO reference were shown in Supplementary Fig. 5b and S5c. EXAFS results suggested the Co was highly metallic with small amounts of oxidized Co. No alloy formation with W was found by fitting the Fourier transform of the first shell (Fig. 3f). The NPs sizes were estimated based on their CN 55 and the results suggested the major phase was extremely small Co nanoparticles (NPs).
The further fittings of all Co/W2TiC2 samples showed Co-O bond distance of 1.97 Å. Unlike the typical Co-O bond distance of 2.13 Å in small CoO clusters reference, the much shorter Co-O bond distance observed in our Co/W2TiC2 samples was very similar to previously reported single site Co structure which had Co+ 2 with 4 Co-N(C) at 1.95 Å 56. Fitting of the Co/W2TiC2-500 showed a Co-O CN of 2.1, suggesting about 50% of the Co existing as single site Co+ 2 on the MXene surface and 50% existing as metallic Co NPs. With the fraction of oxidized Co equal to fit CN/CN of the pure species, metallic Co with Co-Co fitted CN of 2.8 should have a true CN of 5.6 and metallic NPs with this CN should have a size of about 1.5 nm. As the reduction temperature increased, the proportion of oxidized Co decreased, resulting in nearly all metallic Co for Co/W2TiC2-700. For Co/W2TiC2-600 sample, the Co-O CN was about 1, indicating 25% single site Co+ 2 and the true Co-Co CN for the rest 75% metallic Co being 4.4 with NP sized at about 1.0 nm. Co/W2TiC2-700 had a Co-Co CN of 3.8, corresponding to the NPs about 0.9 nm. The summary of percentage of single sites, nanoparticles and nanoparticle sizes were listed in Supplementary Table 3. The higher shell Co-(O)-Co peak of Co-oxide nanoparticles was not detectable due to overlap with the Co-Co scattering path. Subtraction of the Co-Co scattering indicated a very small amount Co-(O)-Co consistent few CoO clusters. Combined with the bond distance, the results suggested the Co on W2TiC2 was highly metallic with certain portion existing as single site Co+ 2 with little surface oxidation. Unlike the rapid oxidation of normal small metallic nanoparticles obtained by other synthesis methods, the tiny Co NPs on W2TiC2 were highly resistive to oxidation even when exposed to air. This further confirmed the strong metal-support interaction between Co and W2TiC2, i.e., the electrons being drawn from Co to W2TiC2 making Co electronically “more positive” and less vulnerable to oxidation. As the reduction temperature increased from 500 ℃ to 700 ℃, the Co structure progressively homogenized, changing from a mixture of single sites and metallic NPs (~ 1.5 nm at 500 ℃) to smaller metallic NPs (~ 1 nm at 600 ℃), and ultimately forming well-dispersed sub-1nm Co NPs (~ 0.9 nm at 700 ℃).
Electrochemical HER evaluation of Co/W 2 TiC 2 catalysts. To explore the catalytic activity of the Co/W2TiC2 catalysts, the HER performance was evaluated in an H-cell using the three-electrode system. The HER performance of the same Co/W2TiC2 catalyst (i.e., Co/W2TiC2-700 sample) in acidic, neutral, and alkaline media were first compared. As shown in Supplementary Fig. 6, the current density in 1 M KOH and 0.5 M H2SO4 significantly exceeds that in 1 M PBS, suggesting the catalyst was more suitable for catalyzing hydrogen evolution in alkaline and acidic environments. Under acidic condition, the HER performance gave the smallest overpotential of 133.92 mV at 10 mA/cm2 with Tafel slope of 72.72 mV/dec. However, a significant degradation in the HER performance was observed after the electrode was immersed in the H2SO4 solution for 30 min without applying current, as shown in Supplementary Fig. 7 and S8. TEM images of the catalyst post-reaction in 0.5 M H2SO4 revealed the degradation of MXene structure and the aggregation of Co into larger particles, ranging from 5–10 nm in size. This was because Co trimer/W2TiC2 and Co tetramer/W2TiC2 were more stable than isolated Co/W2TiC2 in the presence of H*, as seen in our DFT-calculated formation energy of Co clusters at different H* coverages in Supplementary Fig. 17a. In contrast, the Co/W2TiC2-700 sample exhibited no structural damage or performance degradation after the reaction under alkaline conditions, as displayed in Supplementary Fig. 9a. Given the current industrial demand for alkaline water electrolysis catalysts and the observed stability of our Co/W2TiC2-700 sample for alkaline HER, a detailed analysis was conducted to determine the electrocatalytic performance of all the Co/W2TiC2 catalysts under alkaline conditions.
To evaluate the contribution of Co metal and W2TiC2 MXene to the catalytic performances, control experiments were conducted using pure W2TiC2 MXene and Co/WC (commercial tungsten carbide) reduced at 700 ℃. The metal loading on the catalysts was confirmed by inductively coupled plasma optical emission spectrometry (ICP-OES), the Co contents in Co/W2TiC2-500, 600, and 700 were determined to be 4.1%, 4.0%, and 4.3%, respectively, with an approximate average atomic ratio of 4.13%, confirming the total Co amounts in the catalysts are close with no obvious variation across all the three reduction temperatures (Supplementary Table 4). To provide a fair comparison, uniform Co and overall catalyst loading mass were applied on each electrode, as detailed in Supplementary Table 5. All the Co-containing catalysts (Figs. 4a and 4b) exhibited higher HER activities compared to the pristine W2TiC2, underscoring the critical role of Co in the HER activity. Furthermore, all three Co/W2TiC2 catalysts showed superior activity of below 100 mV overpotential compared to Co/WC, which exhibited a 211.93 mV overpotential at current density of 10 mA/cm2 (η @10 mA/cm2). This performance difference could be attributed to the two-dimensional nature of W2TiC2, providing more active sites than the bulk WC. Noticeably, Co/W2TiC2-700 stands out as the best catalyst, with the smallest overpotential of 62.58 mV at 10 mA/cm2, followed by Co/W2TiC2-600 (88.78 mV at 10 mA/cm2) and Co/W2TiC2-500 (99.34 mV at 10 mA/cm2). The relatively subdued HER performance of Co/W2TiC2-500 among the three catalysts can be attributed to the lower fraction and slightly larger metallic NP size among all the three Co/W2TiC2 samples, as explained in the following theoretical modeling section. Notably, the advantage of small overpotential became even more pronounced at higher current densities. Co/W2TiC2-700 showed overpotential of 191.81 mV @ 100 mA/cm², outperforming commercial 40% Pt/C which had an overpotential 238.02 mV @ 100 mA/cm² by 46.21 mV. At higher current density, Co/W2TiC2-700 continued to exhibit excellent low overpotentials as 330.43 mV @ 500 mA/cm² and 407.72 mV @ 1000 mA/cm², highlighting its great potential in industrial-level high current density alkaline hydrogen production.
The Tafel slopes (Fig. 4d) of Co/W2TiC2-700, Co/W2TiC2-600, Co/W2TiC2-500, Co/WC and pristine W2TiC2 were 44.34, 50.78, 71.00, 166.30, and 268.04 mV/dec, respectively. The smallest slope of Co/W2TiC2-700 showed rapid change of current density with slight increases in applied voltage, indicates fast reaction kinetics. The small Tafel slope of Co/W2TiC2-700 also outperformed most of reported non-noble metal catalysts within the realm of alkaline HER, as compared in Supplementary Table 9. The electrochemical impedance spectra (EIS) were collected to study the charge transfer efficiency (Fig. 4e). Nyquist plots showed close solution resistance (Rs) values for all the tests and distinguished charge transfer resistance (Rct) of the catalysts, suggesting similar electrolyte solution environment. Co/W2TiC2-700 and Co/W2TiC2-600 exhibited small semi-circles of the curve, indicating faster charge transfer property. The reduced charge transfer resistance of the Co/W2TiC2 catalysts in comparison to Co/WC and pristine W2TiC2 demonstrated the optimized electron transfer facilitated by the synergistic metal-support interaction between Co and W2TiC2, which agreed with the diminished surface oxidation at higher annealing temperature as observed from XPS and EXAFS analyses.
To further understand the performance of Co/W2TiC2 catalysts, the electrochemical active area (ECSA) was investigated by measuring the double layer capacitance (Cdl). Cyclic voltammograms (CV) were collected at 10 to 100 mV/s scan rates in a non-faradaic region (Supplementary Figs. 12 and 13), and the Cdl values were calculated from the slope of charging current density differences plotted against the scan rates plots. ECSA was determined by dividing the Cdl by the specific capacitance (Cs, assumed as 40 µF/cm2). Supplementary Fig. 12 showed the Cdl of Co/WC and W2TiC2 were 10.43 and 20.56 mF/cm2. The higher Cdl of W2TiC2 was attributable to its inherently larger specific surface area, the characteristic nature of two-dimensional structures. Supplementary Fig. 13 showed Cdl of Co/W2TiC2-500, 600 and 700 are 11.76, 38.61 and 55.47 mF/cm2, respectively. The decrease in Cdl values on Co/W2TiC2 compared with pure W2TiC2 was due to the interaction between the MXene flakes and metal precursors (Co here), as also observed in previous study 42. The high Cdl value of Co/W2TiC2-700 indicated that it had the largest electrochemical active area among all the three catalysts, which could be due to smaller Co nanoparticle size and relatively larger numbers of Co nanoparticle for a constant Co loading in our Co/W2TiC2-700 sample (as evident from Supplementary Table 3). This results also hinted the active sites of Co/W2TiC2 were either Co NPs or the interfaces between Co NPs and W2TiC2 supports, which will be discussed in the next section.
Effect of Co coordination environment on the HER activity and H* binding. Previous literature suggested that the adsorption free energy of H* (\({G}_{H*}\)) was a good descriptor of the HER activity on transition metal electrodes, finding that Pt(111) lies closest to the observed activity peak 57, 58. We therefore calculated relative adsorption free energies (\({\Delta }{G}_{rel}\)) for H* on different Co-modified W2TiC2 materials, using the corresponding value on Pt(111) as a reference: \({\varDelta G}_{rel}={G}_{H*@\text{C}\text{o}/{W}_{2}Ti{C}_{2}}-{G}_{H*@\text{P}\text{t}\left(111\right)}={G}_{H*}+0.21 \text{e}\text{V}\). Smaller magnitudes of \({\varDelta G}_{rel}\) indicate H* binding closer to the previously-observed HER activity peak on pure Pt. The \({\varDelta G}_{rel}\) values calculated based on the most stable binding site of H* on the respective surfaces suggested that the theoretically expected trend in the HER activity was: isolated Co/W2TiC2 (-0.13 eV) > W2TiC2 (-0.32 eV) > Co NP/W2TiC2 (-0.40 eV). This was contrary to our experimental data (Supplementary Table 7) showing a lower overpotential on Co NP/W2TiC2 than on clean W2TiC2. Experimentally, the overpotential for W2TiC2 shifted by -0.24 V relative to Pt, which is in good agreement with the DFT-calculated \({\varDelta G}_{rel}\) on the clean W2TiC2 surface (-0.32 eV). We therefore ascribe the deviation in trend to the strong binding found on the Co NP model: the most stable H-binding site on Co NPs is poisoned by strong binding of H, thereby rendering such sites inactive. We considered \({G}_{H*}\) at a number of possible binding sites in our theoretical model for Co NP/W2TiC2 (Fig. 5a). These suggested that the \({{\Delta }\text{G}}_{rel}\) value (0.02 eV) for H* binding at the Co-W interface was the closest to 0 and therefore is the likely active site. Further, this value is in good agreement with our experimental data showing a shift of + 0.05 V in overpotential relative to Pt. Accounting for these interfacial sites as most active, the theoretically expected HER activity (based on \({\varDelta G}_{rel}\), eV) should be: Co NP/W2TiC2 (0.02) > isolated Co/W2TiC2 (-0.13) > W2TiC2 (-0.32), agreeing with our experimental predictions.
We find that the d-band center of the interfacial W atom acts as a strong descriptor of \({G}_{H*}\) across a series of models, as shown in Fig. 5b, suggesting that Co plays an important role in tuning the electronic structure of W toward favorable interactions with H*. Further discussion on the effect of Co-Co coordination environment and H* surface coverage on the H* binding strength is provided in Supplementary Note 2.
Co/W 2 TiC 2 alkaline HER catalyst under industrially relevant conditions. Stability of the catalyst under reaction environment is one of the crucial factors in evaluating the performance of a HER catalyst for industrial-scale applications. To this end, the stability of Co/W2TiC2-700 was assessed by stepwise and continuous chronopotentiometry tests, as shown in Fig. 4c and 4f. In a 12-hour test, when the current densities were sequentially increased from − 10 to -50 mA/cm2 in the intervals of 4h, the potential remained stable at each setting, and quickly returned to the original values when the current density was reduced from − 50 to -10 mA/cm². Continuous operation at -10 mA/cm2 showed no decrease in potential even after 500 h continuous operation. Moreover, chronopotentiometry test was further conducted under high current density of 1000 mA/cm2, no obvious decay was found within 100 h. In contrast to the rapid degradation of MXene under acidic condition, SEM image of Co/W2TiC2-700 after the stability test in 1M KOH (Supplementary Fig. 9) showed no change in morphology, highlighting the superior stability of Co/W2TiC2-700 under alkaline catalytic condition.
From the temperature-induced structure modulation and metal-support interaction as well as meticulous mechanistic investigations, Co/W2TiC2-700C demonstrated its great potential and scientific grounding to serve as a high-performance water electrolysis catalyst. Considering the practical industrial environment applications that require high current density and the low overpotential exhibited by our Co/W2TiC2-700 at 1000 mA in linear sweep voltammetry, this optimized material was applied in a 2 × 2 cm² flow cell electrolyzer for alkaline water electrolysis tests and hydrogen quantification. Co/W2TiC2-700 was employed as the cathode, paired with non-noble metal material Ni foam as the anode on which oxygen evolution happened, and 1M KOH was used as electrolyte for both sides. Using a potentiostat with a ± 2A/±30V booster, a maximum current of 1.6A (400 mA/cm²) was applied. As shown in Fig. 6b, one piece of Co/W2TiC2-700 electrode was continuously used in chronopotentiometry tests at current densities of 50, 100, 200, 300, and 400 mA/m² for 3 hours each, exhibiting stable performance throughout the 15 hours of operation. The overall cell voltages of Co/W2TiC2-700 working as the cathode from 50 to 400 mA/cm² were all lower than that of Pt/C with comparable loading amount (0.57 mg Pt/cm²) on the identical carbon paper substrate. The lower cell voltage under the same current indicated less voltage was required to drive the electrolysis reaction beyond its thermodynamic potential, resulting from the synergistic effects of intrinsic catalytic activity, conductivity, and surface property. The hydrogen produced was quantified using on-line gas chromatography (GC), as shown in Fig. 6c. The Co/W2TiC2-700 catalyst achieved nearly 100% H2 faradaic efficiency and comparable hydrogen production rates as commercial Pt/C at the conditions of 50 to 400 mA/cm², with the highest hydrogen production rate of 706.75 mL/h attained at 1.6 A. Furthermore, to verify the possibility of further reducing the cell voltage, IrO2 anode and high operation temperature of 60 ℃ were applied to the same flow cell, and cell voltage was further reduced to 2.85 mV at 400 mA/cm² (Supplementary Fig. 21). Energy efficiency analysis is provided in Supplementary Note 3, the IrO2||Co/W2TiC2-700 demonstrated 46.98%-65.33% iR-free cell energy efficiency, higher than 44.63%-58.31% compared to Pt, substantiated its feasibility for efficient hydrogen production at practical high current applications.