DOI:
10.1039/D4TA03999H
(Paper)
J. Mater. Chem. A, 2024, Advance Article
Synthesis of two-dimensional N-terminated molybdenum carbides using an alloying strategy in molten salt†
Received
10th June 2024
, Accepted 1st August 2024
First published on 6th August 2024
Abstract
Molybdenum carbide (Mo2CTx) has an electronic structure similar to that of platinum, thereby making it a promising non-precious-metal catalyst. However, preparing Mo2CTx efficiently and rapidly using a conventional redox method is challenging owing to the high exfoliation energy of the Mo2Ga2C precursor. In this study, a novel alloying method was developed to etch Ga rapidly from its precursor. The etching process was conducted in a LiCl–KCl molten salt containing Li3N and an iron triad element (Fe, Co, or Ni). Here, Ga formed an alloy with the iron triad element, thereby reducing the reaction energy barrier and promoting the kinetics of Ga etching. Simultaneously, adding Li3N to the molten salt created a specific N ion as a stable functional group adsorbed on the surface of Mo2C. This nitrogen functional group (–N) optimizes the electronic structure of surface molybdenum by inducing strong covalent bonds. Consequently, the synthesized N-terminated Mo2CTx demonstrates superior electrocatalytic oxygen evolution activity to conventional F-terminated Mo2CTx and commercial IrO2 in a 1 M KOH solution.
1 Introduction
MXenes are two-dimensional materials like transition metal carbides, nitrides, and carbonitrides.1–5 Owing to their distinctive structures, electronic properties, and mechanical characteristics, they have garnered considerable attention in various fields.6–8 Ti3C2Tx, the most widely used MXene, is used as a supercapacitor.9–13 Its preparation generally involves a redox method wherein the Al element is etched from the ternary Ti3AlC2 precursor using hydrofluoric acid (HF) or a Lewis acid molten salt as the oxidation agent.8,14–16
Unlike the conventional MXene, which is widely used as a supercapacitor, Mo2CTx has an electronic structure similar to that of platinum, thereby making it suitable for various catalytic applications.17–19 However, unlike the efficient Al etching from the Ti3AlC2 precursor, rapid and efficient Ga etching is a considerable challenge owing to the higher exfoliation energy of the Mo2Ga2C precursor.8 Synthesis of Mo2CTx reportedly required nearly a week of Ga etching in highly oxidative HF or LiF-HCl solutions.20–22 To date, Mo2CTx with surface functional groups other than –F, –Cl, and –OH has not been synthesized in a single step, despite the potential to chemically modify surface terminations (Tx) in MXenes by adjusting the etching agent or through post-treatment.23
Our study was inspired by the observation that interactions between the external metals and MAX phases can induce the diffusion of A-site atoms, thereby forming new phases. For example, Au reacts with Ti3SiC2 to form Ti3AuC2 and Ti3Au2C2,24 while Cu reacts with Ti3AlC2 to form Ti3(Al1−δCuδ)C2.25 By leveraging atomic diffusion to reduce the reaction energy barrier, we can synthesize new MAX phases that are typically challenging to obtain using conventional solid sintering methods.
In this study, we propose a novel approach for efficient Ga etching from a Mo2Ga2C precursor to produce Mo2CTx, which involves using an iron triad element (Fe, Co, or Ni) and Li3N in a KCl–LiCl molten salt. The iron triad element forms an alloy with Ga, thereby lowering the reaction energy barrier and facilitating Ga etching. Simultaneously, adding Li3N to the molten salt introduces new nitrogen functional groups and promotes the formation of Mo2CN2 nanoflakes. The synthesized N-terminated Mo2CTx exhibits higher electrocatalytic oxygen evolution reaction activity than Mo2CTx prepared using conventional HF solution and commercial IrO2 in a 1 M KOH solution. This new strategy not only enables the one-step preparation of novel MXenes, but also presents new possibilities for their application in electrocatalysis.
2 Experimental section
2.1 Chemicals and materials
Molybdenum gallium carbon (Mo2Ga2C, 99 wt%) was purchased from 11 technology company (China). Lithium nitride (Li3N, 99.4 wt%), potassium chloride (KCl, 99.8 wt%), lithium chloride (LiCl, 99.99 wt%), nickel powder (Ni, 99.5 wt%), cobalt powder (Co, 99.9 wt%), iron powder (Fe, 99.9 wt%), Nafion (5 wt%), isopropyl alcohol (99.7 wt%), potassium hydroxide (KOH, ≥85.0 wt%), iridium oxide (IrO2, 99 wt%), acetylene black (99.9 wt%), hydrochloric acid (HCl, 37 wt%), and hydrogen fluoride (HF, 40 wt%) were purchased from Sinopharm Chemical Reagent Co., Ltd.
2.2 Synthesis of A-MS-Mo2CTx
Mo2Ga2C (100 mg), metal powder (Ni or Co or Fe, 100 mg), and Li3N (200 mg) were thoroughly mixed and pressed into a tablet at 30 MPa under an air atmosphere. The prepared tablet was placed in a glassy carbon crucible and transferred to a resistance furnace within an argon-protected glovebox. The reaction temperature was increased from room temperature to 700 °C at a rate of 3 °C min−1 and maintained at 700 °C for 24 h. After naturally cooling to room temperature, the product was washed with deionized water and HCl to remove residual salt and metal impurities, respectively. Finally, A-MS-Mo2CTx was obtained by drying in a vacuum oven for 12 h.
2.3 Synthesis of HF-Mo2CTx
LiF (1.0 g) was added to 15 mL of 12 M HCl and stirred continuously in a Teflon reactor at 90 °C for 1 h. After gradually adding 0.5 g of Mo2Ga2C and 5 mL of HF solution, the Teflon reactor was kept in a muffle furnace at 150 °C for 72 h and naturally cooled to room temperature.26 The product was centrifuged at 5000 rpm for 5 min, and rinsed using deionized water, and this was repeated several times until the pH of the supernatant reached 6–7. Finally, the product was dried at 80 °C for 24 h in a vacuum oven to obtain HF-Mo2CTx.
2.4 Material characterization
X-ray diffraction (XRD) analyses were performed using a Bruker D8 Advance X-ray diffractometer with Cu-Kα radiation (λ = 1.54 Å). Morphological analyses were conducted using transmission electron microscopy (TEM, JEM-2100) coupled with energy-dispersive X-ray spectroscopy (EDX) and scanning electron microscopy (SEM, FEI Apreo S) coupled with energy-dispersive spectroscopy (EDS). X-ray photoelectron spectroscopy (XPS) was performed on an ESCALAB 250Xi (Thermo Fisher) spectrometer using an Al-Kα radiation (λ = 8.34 Å) power source. Peak fitting was performed using Casa XPS. Raman spectroscopy data were collected on a Lab RAMHR800 instrument (Jobin Yvon, France) equipped with an air-cooled CCD array detector in the backscattering configuration. The 473 nm line of an argon ion laser (100 mW) was used to excite the sample. The atomic structure was analyzed using transmission electron microscopy (STEM) mode on a double Cs corrected FEI Titan Themis Z60-300 system equipped with a Super-X EDS, operated at 300 kV. The Mo K-edge X-ray absorption spectroscopy (XAS) data were collected at the 14W1 beamline in the Shanghai Synchrotron Radiation Facility (SSRF). The X-rays were monochromatized using a double-crystal Si (111) monochromator, and the energy was calibrated using molybdenum metal foil for the Mo K-edge. A gas chromatograph (GC2060) equipped with three thermal conductivity detectors was used to determine the gas composition during the reaction process.
2.5 Evaluation of catalytic performance
Electrochemical tests were conducted using an Autolab electrochemical workstation (PGSTAT 302 N, Metrohm Auto Lab Co., Ltd) with a rotating-disk three-electrode system (RDE, Pine Research Instrumentation). Prior to each test, 1 M KOH was saturated with O2 for 30 min at room temperature (approximately 25 °C). A mixture of isopropyl alcohol and deionized water was prepared in a 7:3 volume ratio. Furthermore, 10 mg of catalyst, 5 mg of acetylene black, and 80 μL of Nafion (Sigma-Aldrich) were added and ultrasonicated for 0.5 h to form a uniform catalyst ink. The prepared ink was later applied to a glassy carbon (GC) electrode with an area of 0.196 cm2 to create a working electrode. A graphite rod and Hg/HgO (1.0 M KOH) electrodes served as the counter and reference electrodes, respectively. The potentials versus Hg/HgO were converted to the reversible hydrogen electrode (RHE) as expressed in eqn (1) as follows:27,28 |
ERHE = EHg/HgO + 0.0591 × pH + 0.097
| (1) |
LSV polarization curves were measured in a 1.0 M KOH aqueous electrolyte using a rotating-disk electrode at 1600 rpm with a scan rate of 5 mV s−1. Tafel slopes were calculated using eqn (2) by fitting the linear portions of the LSV curve at a low overpotential as follows:29
|
η = blogj + a
| (2) |
where
b is the Tafel slope,
η the overpotential, and
j the current density.
Electrochemical impedance spectroscopy (EIS) analysis was conducted over a frequency range from 100 kHz to 100 mHz at a current density of 10 mA cm−2. Cyclic voltammetry (CV) curves were recorded within the potential range of 0.1–0.3 V vs. RHE at various scan rates. The Cdl was calculated by plotting the current differences at a potential of 0.225 V vs. RHE against the scan rates as expressed in eqn (3) as follows:30
where
Ic represents the charging current and
v the scan rate.
2.6 Computation details
Density functional theory (DFT) calculations5 were conducted using a frozen-core projector-augmented wave (PAW) method31 with the Vienna ab initio simulation package (VASP).32 The Perdue–Burke–Ernzerhof (PBE) version of the generalized gradient approximation (GGA) was used to treat the electron–electron exchange and correlation interactions.33 Transition-state calculations were conducted using a CI-NEB method.34 DFT was used to analyze the process of etching Ga from Mo2Ga2C and to assess the stability of the surface functional groups. The transition states for Ga etching from Mo2Ga2C via alloying and oxidation were simulated using the nudged elastic band (NEB) method. The K-point was set to 3 × 6 × 1, while the vacuum layer and plane wave cut-off energy were 15 Å and 500 eV, respectively. Owing to the limitations of the super-soft pseudopotential in reciprocal space, which affects exchange-correlation and underestimates the band gap for GGA functions, a GGA + U method was employed to accurately describe the d orbitals of transition metal elements. The Ueff values of Mo, Ga, and Ni are 4.0,35 3,36 and 2.95 eV,37 respectively. The atomic positions were optimized until the total energy change was less than 1 × 10−6 eV per atom and maximum force on each atom was less than 0.01 eV Å−1. As regards the calculation of adsorption energy of functional groups with different valence states, varying numbers of Li atoms were placed and fixed approximately 4 Å above the surface of the Mo2Ga2C supercell to adjust the valence of the terminal groups. The adsorption energy on the surface was expressed using eqn (4) as follows:38 |
Eads = ET/slab − Eslab − ET
| (4) |
where ET/slab and Eslab represent the energies with and without the termination adsorbate, respectively, and ET the energy of the termination adsorbate in an empty box. Here, the classical d-band models can be used as effective descriptors to evaluate electrocatalytic reactions.39 The band center level (b) is expressed using eqn (5) as follows:40 |
| (5) |
where D (E) represents the density of states (DOS) of Mo d-orbitals. Here, the model identifies the position of the orbital center and filling degree of the orbitals. Furthermore, the d-band center level shifts down deeper than the Fermi level, thereby resulting in higher d-electron occupation in σ* states.41
3 Results and discussion
3.1 Structure of Mo2CTx etched using the alloying method in molten salt
Mo2CTx was synthesized using Mo2Ga2C as a precursor via a sequence of alloying and etching steps, as shown in Fig. 1a. The Mo2Ga2C was mixed with Ni (Fe or Co) powder, Li3N, and a eutectic KCl–LiCl molten salt, and was later pressed into a pellet shape (Fig. S1†). The pellet was heated at 700 °C for 24 h in an inert atmosphere. The product was washed using deionized water and hydrochloride to remove residual molten salts and GaNi3 alloy byproducts (Fig. S2†). This product was labeled as A-MS-Mo2CTx. As shown in the XRD pattern in Fig. 1b, the characteristic peaks of Mo2Ga2C were observed at 9.93°, 34.19°, 37.65°, 39.96°, 42.85°, 49.61°, 53.46°, and 61.17°, which correspond to the (002), (100), (103), (008), (105), (107), (108), and (110) planes, respectively. These peaks correspond with those reported in the literature.42 After alloying and etching in the molten salt, the distinct peaks corresponding to Mo2Ga2C considerably reduced, while five new characteristic peaks emerged at 6.46°, 12.87°, 19.40°, 25.81°, and 32.44°, thereby indicating a notable increase in the c-lattice parameter (c-LP) from 17.86 to 26.60 Å. Contrary to the Mo2Ga2C precursor, which had a high Ga content and closely stacked layers (Fig. S3†), A-MS-Mo2CTx showed a structure comprising ultra-thin sheets arranged in a flower-like pattern (Fig. 1c). The Ga content in the product was considerably reduced from 13.06 to 1.00% (Fig. S3 and S4†). Adjusting the reaction temperature and time further decreased the Ga content (Fig. S5 and S6†). The uniform distribution of Mo, C, and N on the surface (Fig. 1d) indicates that –N functional groups were successfully introduced into A-MS-Mo2CTx. Transmission electron microscopy (TEM) images showed that A-MS-Mo2CTx had an ultra-thin nanosheet layer morphology (Fig. 1e) with an interlayer spacing of 1.37 nm (Fig. 1f), which corresponds to the XRD analysis. STEM images confirmed the lamellar microstructure of A-MS-Mo2CTx (Fig. S7†). Furthermore, electron energy loss spectroscopy (EELS) detected an N K-edge at 400 eV (Fig. 1g), thereby indicating the incorporation of N into A-MS-Mo2CTx. EDS and EDX analyses determined a Mo/N ratio of approximately 1 (Fig. S4 and S8†), thereby indicating the formation of Mo2CN2 following its etching in the Ni–Li3N–KCl–LiCl system. Additionally, the content of the N groups on the surface can be adjusted by varying the ratio of Mo2Ga2C to Li3N (Fig. S9†).
|
| Fig. 1 Synthesis procedure and characterization of A-MS-Mo2CTx: (a) schematic of A-MS-Mo2CTx prepared by Ga etching in molten salt, (b) XRD patterns of Mo2Ga2C (blue line) and A-MS-Mo2CTx (red line), (c) SEM image of the A-MS-Mo2CTx nanosheet, (d) EDS mapping images of Mo, C, and N elements, (e) TEM image of the A-MS-Mo2CTx nanosheet, (f) high resolution transmission electron microscopy (HRTEM) image of the A-MS-Mo2CTx nanosheet, and (g) EELS spectrum of the N K-edge in the A-MS-Mo2CTx nanosheet. | |
To further investigate the surface structure of A-MS-Mo2CTx, numerous techniques were used, including X-ray photoelectron spectroscopy (XPS), Raman spectroscopy, X-ray absorption near-edge structure (XANES), and extended X-ray absorption fine structure (EXAFS). The distinctive peaks of Ga, including Ga 2p (binding energies of 1144.7 and 1117.9 eV),18,43 Ga 3s, Ga 3p, and Ga 3d (centered at 160.5, 105.5, and 19.2 eV, respectively) were considerably reduced after Mo2Ga2C was etched by the alloying process in molten salt (Fig. 2a). As presented in Table S1†, the Ga 2p content decreased significantly from 17.56% in Mo2Ga2C to 1.63% in A-MS-Mo2CTx, thereby indicating that a considerable amount of the Ga had been eliminated.42
|
| Fig. 2 Structural characterization of A-MS-Mo2CTx: (a) survey XPS spectra of Mo2Ga2C and A-MS-Mo2CTx, (b) high-resolution XPS Mo 3d spectra of A-MS-Mo2CTx and Mo2Ga2C, (c) high-resolution XPS C 1s spectra of A-MS-Mo2CTx and Mo2Ga2C, (d) N 1s spectrum of A-MS-Mo2CTx, (e) Raman spectra of A-MS-Mo2CTx and HF-Mo2CTx, (f) normalized XANES spectra of the Mo K-edge for Mo foil, MoO2 powder, MoO3 powder, Mo2Ga2C, and A-MS-Mo2CTx, (g) calculated chemical valences for Mo in Mo2Ga2C and A-MS-Mo2CTx, (h) FT-EXAFS spectra of the Mo K-edge for A-MS-Mo2CTx, Mo2Ga2C, Mo foil, and MoO2, and (i) wavelet transform (WT) of A-MS-Mo2CTx and Mo2Ga2C. | |
The high-resolution XPS spectra of Mo2Ga2C at the Mo 3d core level exhibited three doublet peaks at 227.8 (230.9 eV), 228.8 (232.0 eV), and 232.2 eV (235.4 eV), corresponding to Mo–C, Mo4+, and Mo5+, respectively (Fig. 2b and Table S2†). The Mo valence states (+4 and +5) were attributed to their surface oxides.42,44 Conversely, the Mo 3d XPS spectra for A-MS-Mo2CTx exhibited only two doublet peaks at 229.2 (232.4) eV and 232.0 (235.1) eV, corresponding to N–Mo–C and Mo5+,18,44 respectively (Fig. 2b and Table S2†). This shift toward higher valence states of Mo in A-MS-Mo2CTx can be attributed to the removal of Ga and introduction of N-termination. The C 1s spectrum of Mo2Ga2C exhibited four peaks at 283.3, 284.8, 285.7, and 288.7 eV (Fig. 2c and Table S3†), which correspond to the Mo–C, C–C, C–O, and CO bonds, respectively.18,42 Contrarily, the binding energy of the Mo–C bond shifted toward a higher value in the A-MS-Mo2CTx spectrum. Additionally, a new peak at 286.5 eV was identified, thereby corresponding to the C–N bond (Fig. 2c).44 The N 1s high-resolution spectrum further showed that the main peak at 397.2 eV corresponded to the Mo–N bond, while the peaks at 398.0, 399.2 and 400.3 eV were attributed to pyridine N, pyrrole N, and graphitic N, respectively (Fig. 2d and Table S4†).44–46 The ratio of Mo/N determined using XPS was approximately 1 (Table S1†), which corresponds to the findings from STEM and EDS analyses, further supporting the formation of Mo2CN2. Raman spectroscopy was used to analyze Mo2CTx with various functional groups. As shown in Fig. 2e, Mo2CTx synthesized using HF (HF-Mo2CTx, Fig. S10†) displayed two peaks at 91 and 285 cm−1, which correspond to the Eg and Ag modes of the MXene flakes dominated by Mo atoms, respectively.47 The presence of F and O terminated groups derived from the HF solution, which were randomly distributed at different sites on the surface of the Mo atoms, resulted in complex Raman peaks of HF-Mo2CTx (labeled as *). However, the characteristic peak of Mo–C in A-MS-Mo2CTx was enhanced at 88 cm−1 owing to the introduction of the N-terminations on the surface, thereby indicating that A-MS-Mo2CTx had a more ordered Mo–C framework and better crystallinity, which is in agreement with the XRD results. The in-plane A1g band at 660 cm−1 represented the vibrations of the Mo atoms on the surface, with its position and width influenced by the atoms attached to Mo.47,48 The peak positions differed between HF-Mo2CTx (664 cm−1) and A-MS-Mo2CTx (632 cm−1) owing to the different surface functional groups. Further investigation into the local atomic coordination and electronic structure of Mo was conducted using XAS. The absorption edge positions of 20006.7, 20012.9, and 20016.0 eV corresponded to various Mo valence states in Mo (0), MoO2 (+4), and MoO3 (+6), respectively (Fig. 2f). The absorption edges of Mo in Mo2Ga2C and A-MS-Mo2CTx were measured at 20009.6 and 20012.1 eV, respectively. The XANES results indicated that the Mo valence states in Mo2Ga2C and A-MS-Mo2CTx were approximately +2 and +3.5, respectively (Fig. 2g), thereby indicating that the removal of Ga atoms and introduction of N-terminations increased the average Mo valence state in A-MS-Mo2CTx, which corresponded to the XPS analysis. The Fourier-transform EXAFS spectra revealed peaks at 1.63 and 2.63 (2.23) Å relative to the coordination of Mo–N/C (Mo–Tx) and the Mo–Mo shell for A-MS-Mo2CTx in R-space, respectively (Fig. 2h). The fitting bond length of Mo–N/C (Mo–Tx/C) was 2.11 Å, while the two Mo–Mo bond lengths were 2.79 and 3.10 Å, thereby corresponding to the Mo–Mo distance between different layers and within a similar layer, respectively (Fig. S11 and Table S5†). The first shell scattering coordination number for Mo–N/C (Mo–Tx/C) was 5.9, which is considerably higher than that of Mo2Ga2C (3.0 as presented in Table S5†). Wavelet transform (WT) analysis of the Mo K-edge EXAFS oscillations showed that the maximum WT for A-MS-Mo2CTx at 5.0 and 10.0 Å−1 was attributed to the Mo–C/Tx and Mo–Mo bonds (Fig. 2i), respectively, which is different from those of Mo2Ga2C (4.0 and 8.0 Å−1, Fig. 2i). These XAS results further confirmed the removal of Ga and introduction of N surface terminations.
3.2 Experimental and density-functional theory (DFT) analyses of the Mo2Ga2C etching mechanism
The formation of MXenes involve the etching of element A from the MAX precursor and adsorption of functional groups.13,49,50 The etching process generally alters the valence state of element A via conventional redox reactions.8,51,52 However, in the etching product GaNi3 alloy (Fig. S2 and S12†), the valence state of Ga remains unchanged compared to its state in the Mo2Ga2C precursor †. Hence, two possible reaction pathways may result in the formation of GaNi3 alloys. One pathway is the conventional redox process, wherein Ga in Mo2Ga2C is initially oxidized to Ga2O3 in molten salt (Fig. S14†), which is further reduced to Ga by Li3N, and finally alloyed with Ga to form GaNi3. Another etching pathway involves the direct alloying of Ga in Mo2Ga2C with Ni, thereby forming GaNi3 alloy particles without any redox reaction. To further investigate the Ga etching pathways, DFT calculations were performed. The etched products of Ga2O3 and GaNi3 were modeled above the (110) interface of Mo2Ga2C, both featuring Ga defect sites (Fig. S15 and S16†). As shown in Fig. 3, Ga migrates from the Mo2Ga2C interlayer to the defect sites, thereby forming Ga2O3 via oxidation or GaNi3 via alloying. The Gibbs free energy of Ga2O3 formation is −3.4 eV, which is considerably lower than −0.28 eV for GaNi3 alloy formation, thereby indicating that Ga is thermodynamically more likely to be etched via oxidation. However, transition state searches using the climbing image nudged elastic band (CI-NEB) method (Fig. S17 and S18†) revealed that the kinetic energy barrier for Ga oxidation is 1.5 eV, which is higher than the 1.18 eV for Ga alloying. Hence, a direct alloying pathway is more favorable for dynamic Ga etching from the Mo2Ga2C interlayer. Furthermore, in the presence of Ni and absence of Li3N in the reaction system, Ga atoms can also be etched via alloying to form GaNi3 (Fig. S19†), thereby confirming that Ga etching from Mo2Ga2C occurs via a direct alloying path.
|
| Fig. 3 Reaction coordinates of A-MS-Mo2CTx. | |
Although Ga atoms can be etched by alloying with Ni, the etched Mo2C* structure is unstable without Li3N in the reaction system and will transform into β-Mo2C at high temperatures (Fig. S19†). Hence, the adsorption of anions from molten salt to form functional groups is crucial for maintaining a stable layered structure.53 Considering that Li3N easily hydrolyzes to LiOH,54,55 the adsorption energies of different anions, such as N, Cl, OH, and O as surface termination were calculated. The adsorption energy of N generated by the LiN cluster as the functional groups is −3.67 eV (Fig. 3), which is considerably lower than those of cluster LiCl, Li2O, and LiOH. Additionally, as the Ga atoms are etched and number of functional groups increases, the adsorption energy of the LiN cluster on the Mo2C* surface decreases (Fig. S20–S22†). Hence, more N anions produced by the LiN cluster can be easily adsorbed on the Mo surface, thereby forming a more stable phase (Mo2CN2) than that the other anions owing to the lower average adsorption energy (Fig. S23†). Furthermore, H2 was detected using gas chromatography when Mo2Ga2C was etched in the Ni–Li3N–KCl–LiCl system (Fig. S24†), thereby indicating that LiOH produced by Li3N hydrolysis was involved in the reaction. Hence, the etching reaction mechanism can be expressed using eqn (6). Herein, the Ga atoms in Mo2Ga2C are etched by forming GaNi3. Simultaneously, the specific N ions produced by the reaction of Li3N and LiOH are adsorbed on the surface of Mo2C* as a stable functional group, expressed as follows:
|
Mo2Ga2C + 6Ni + 2Li3N + 6LiOH → Mo2CN2 + 2GaNi3 + 6Li2O + 3H2↑
| (6) |
As a synthetic medium, an appropriate amount of LiCl–KCl can improve the homogeneity of the reaction and promote the conversion of Mo2Ga2C to Mo2CN2 (Fig. S25†). Under similar conditions, Mo2Ga2C can be etched by alloying with Fe or Co in a KCl–LiCl molten salt containing Li3N, thereby resulting in the formation of Mo2CN2 (Fig. S26 and S27†).
3.3 Improvement of electrochemical properties by N-termination
Incorporating non-metal heteroatoms (N, O, and S) into MXenes is effective in regulating the surface chemistry, thereby facilitating electron/ion movement and enhancing the electrocatalytic activity.56 Tang et al.57 found that Ti3C1.6N0.4 and Ti3C1.8N0.2, with a higher N content and pyridinic N species, exhibited higher OER activity than pristine Ti3C2Tx. This implies that Mo–N species can promote charge transfer and improve the catalytic performance. To evaluate the OER activity, linear sweep voltammetry (LSV) was conducted on A-MS-Mo2CTx, HF-Mo2CTx, and commercial IrO2 in 1 M KOH (Fig. 4a). As shown in Fig. 4b, the overpotential at 10 mA cm−2 for A-MS-Mo2CTx was 320 mV, which is considerably lower than that of conventional F-terminated HF-Mo2CTx (570 mV) and commercial IrO2 (410 mV). Furthermore, A-MS-Mo2CTx exhibited a Tafel slope of 53.7 mV dec−1, which surpasses that of HF-Mo2CTx (82.8 mV dec−1) and IrO2 (73.3 mV dec−1), thereby indicating enhanced OER kinetics and a faster reaction rate (Fig. 4c). EIS was used to investigate the charge and mass transfer characteristics of the materials, and the results are shown in Fig. 4d and Table S6.† A-MS-Mo2CTx exhibited a smaller semicircle (8.6 Ω) than IrO2 (58.2 Ω) and HF-Mo2CTx (101 Ω) (Fig. S28 and Table S6†), thereby indicating its lower charge transfer impedance. This reaction impedance can be attributed to the N surface termination in Mo-based MXene, which enhances the migration of electrons and ions. To calculate the electrochemically active surface area (ECSA), the double-layer capacitance (Cdl) was measured using CV at different sweep rates (Fig. S29†). The Cdl of A-MS-Mo2CTx was 18.6 mF cm−2 (Fig. S30†), which corresponds to an estimated active site area of approximately 465 cmECSA2 cmgeo−2.30 Compared to the Cdl of IrO2 (1.9 mF cm−2), the ECSA of A-MS-Mo2CTx was approximately ten times larger than that of IrO2. The ECSA-normalized oxygen evolution current (j) was used to highlight the intrinsic catalytic activity (Fig. S31†). The results indicate that the superior OER performance of N-terminated Mo2CTx, relative to IrO2, is principally attributed to its enhanced ECSA, which is an advantage among two-dimensional materials. Furthermore, the electrocatalytic durability of A-MS-Mo2CTx has been tested at a current density of 10 mA cm−2, wherein it maintained 97% of its potential after 60 h (Fig. 4e). Even at a high current density of 500 mA cm−2, A-MS-Mo2CTx maintained relative stability (Fig. S32†). Consequently, A-MS-Mo2CTx exhibits outstanding OER performance, thereby surpassing most MXene composites (Fig. S33†) and rivaling precious-metal benchmark catalysts.
|
| Fig. 4 Electrochemical measurements in 1 M KOH: (a) LSV curves of different catalysts, (b) overpotentials of catalysts obtained at 10 mA cm−2, (c) Tafel plots of catalysts, (d) EIS tests of catalysts, and (e) durability and stability tests of A-MS-Mo2CTx and IrO2 at a constant current density of 10 mA cm−2. | |
To further understand the OER mechanism, the structures of the samples were characterized using ex situ XRD after different reaction times. As regards HF-Mo2CTx, the characteristic peak at 8.2° gradually disappeared after the OER (Fig. S34†), thereby indicating the destruction of its two-dimensional layer structure using high ECSA, which correspondingly reduced its activity. However, the characteristic peaks at 6.46° and 12.87° for A-MS-Mo2CTx persisted throughout the OER process (Fig. S35†), thereby indicating that the stable two-dimensional layer structure was maintained, resulting in good durability over long reactions. Furthermore, the stability was confirmed by the SEM-EDS results, thereby showing that the morphology of the layered structure remained unchanged, and Mo, C, N and O elements were uniformly distributed (Fig. S36†).
Furthermore, A-MS-Mo2CTx was characterized using XAS and XPS after the OER to explore the origins of its high activity. The Mo K-edge XANES analysis indicated that the valence of Mo increased from +3.5 to +4.0 after the OER (Fig. S37†), which is beneficial for OER activity.23 Considering that A-MS-Mo2CTx maintains a stable 2D structure during the OER, it is speculated that the increase in the Mo valence state is owing to the change in the surface functional group during the adsorption/desorption of intermediate products. Surface electronic information is obtained using XPS, which has a detection depth of 2–5 nm.58 Compared to the bulk, the Mo valence on the surface after the OER considerably increased, with the content of Mo6+ reaching 38.2% (Fig. S38 and Table S7†), which is likely influenced by the change in the surface functional groups. The change in Mo valence in HF-Mo2CTx was also investigated using XPS. The results showed that the Tx–Mo–C bond disappeared, thereby indicating a corresponding decrease in the activity (Fig. S38 and Table S7†). To further understand the difference in Mo valence between A-MS-Mo2CTx and HF-Mo2CTx, DFT calculations were performed. The Mo-based MXene models with –F, –N, and –O surface terminations were constructed to simulate the effects of different functional groups (Fig. S39 and Table S8†). The calculated Bader charges for Mo were found to be +1.55, +1.61, and +1.78, for –F, –N, and –O-termination, respectively, thereby corresponding to experimental observations of the valence states. Beyond the charge distribution at Mo sites, surface functional groups also influence the d-band centers, which are key descriptors for OER activity.39,59 The d-band center of Mo with an N-termination is approximately −2.52 eV, which is lower than that with an O-termination (−1.96 eV) and F-termination (−1.88 eV) (Fig. S40†). This downshift results in a more stable bonding state and stronger covalent interaction,60,61 thereby enhancing the OH− absorption and enabling the OER. Electron localization function (ELF) analysis revealed that N-terminated Mo2C frameworks exhibit weakened ionization around Mo atoms, forming stable N–Mo–C bonds and improving the stability of the 2D structure (Fig. S41†). Hence, the N-terminated Mo2CTx optimizes the electronic structure of surface Mo by inducing high valence and strong covalency, thereby promoting the OER catalytic performance.
4 Conclusions
A novel and highly efficient alloying strategy was developed in this study for synthesizing N-terminated Mo2CTx. Ga atoms in Mo2Ga2C were selectively etched into alloys using an iron triad element (Fe, Co, or Ni) in LiCl–KCl molten salt with Li3N. The N clusters generated by Li3N in the molten salt were adsorbed on the Mo surface as functional groups, thereby forming Mo2CN2 nanoflakes. The direct alloying process reduced the reaction energy barrier and enhanced the etching kinetics. The synthesized N-terminated Mo2CTx exhibited higher electrocatalytic OER activity than conventional HF-Mo2CTx and commercial IrO2 under alkaline conditions. This novel direct alloying method shows great potential for synthesizing a wide variety of 2D materials and paves the way for new explorations of their applications in electrocatalysis.
Data availability
The data that support the findings of this study are available from the corresponding authors, M. Shen & J. Zhou & J.-Q. Wang, upon reasonable request.
Author contributions
Weiyan Jiang: writing – original draft, validation, investigation, formal analysis, data curation. Zihan Gao: data curation, formal analysis, writing – review & editing. Miao Shen: conceptualization, investigation, validation, writing – review & editing, funding acquisition. Jing Zhou: methodology, investigation, formal analysis, writing – review & editing. Rui Tang: validation, software, methodology. Chuangqiang Wu: methodology, formal analysis. Linjuan Zhan: resources & formal analysis. Jian-Qiang Wang: funding acquisition, resources, supervision.
Conflicts of interest
The authors declare that they have no conflict of interest.
Acknowledgements
This study was financially supported in part by the Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDA0400000), Natural Science Foundation of Shanghai (Grant No. 21ZR1476200), Zhejiang Provincial Natural Science Foundation of China (LR24A050001), and “Transformational Technologies for Clean Energy and Demonstration” Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDA21000000). The authors also appreciate the Shanghai Synchrotron Radiation Facility (14W1, SSRF) for their support in material characterization.
References
- H. Ding, Y. Li, M. Li, K. Chen, K. Liang, G. Chen, J. Lu, J. Palisaitis, P. O. Å. Persson, P. Eklund, L. Hultman, S. Du, Z. Chai, Y. Gogotsi and Q. Huang, Science, 2023, 379, 1130–1135 CrossRef CAS PubMed .
- D. Wang, C. Zhou, A. S. Filatov, W. Cho, F. Lagunas, M. Wang, S. Vaikuntanathan, C. Liu, R. F. Klie and D. V. Talapin, Science, 2023, 379, 1242–1247 CrossRef CAS .
- M. Naguib, M. Kurtoglu, V. Presser, J. Lu, J. Niu, M. Heon, L. Hultman, Y. Gogotsi and M. W. Barsoum, Adv. Mater., 2011, 23, 4248–4253 CrossRef CAS .
- M. Naguib, M. W. Barsoum and Y. Gogotsi, Adv. Mater., 2021, 33, 2103393 CrossRef CAS .
- W. Kohn and L. J. Sham, Phys. Rev., 1965, 140, A1133 CrossRef .
- A. VahidMohammadi, J. Rosen and Y. Gogotsi, Science, 2021, 372, eabf1581 CrossRef CAS PubMed .
- X. Li, Z. Huang, C. E. Shuck, G. Liang, Y. Gogotsi and C. Zhi, Nat. Rev. Chem, 2022, 6, 389–404 CrossRef .
- K. R. G. Lim, M. Shekhirev, B. C. Wyatt, B. Anasori, Y. Gogotsi and Z. W. Seh, Nat. Synth., 2022, 1, 601–614 CrossRef .
- X. Wang, T. S. Mathis, K. Li, Z. Lin, L. Vlcek, T. Torita, N. C. Osti, C. Hatter, P. Urbankowski, A. Sarycheva, M. Tyagi, E. Mamontov, P. Simon and Y. Gogotsi, Nat. Energy, 2019, 4, 241–248 CrossRef CAS .
- M. Shen, W. Jiang, K. Liang, S. Zhao, R. Tang, L. Zhang and J.-Q. Wang, Angew. Chem., Int. Ed., 2021, 60, 27013–27018 CrossRef CAS PubMed .
- C. Liu, Y. Bai, W. Li, F. Yang, G. Zhang and H. Pang, Angew. Chem., Int. Ed., 2022, 61, e202116282 CrossRef CAS PubMed .
- M. K. Aslam, Y. Niu and M. Xu, Adv. Energy Mater., 2021, 11, 2000681 CrossRef CAS .
- M. R. Lukatskaya, O. Mashtalir, C. E. Ren, Y. Dall'Agnese, P. Rozier, P. L. Taberna, M. Naguib, P. Simon, M. W. Barsoum and Y. Gogotsi, Science, 2013, 341, 1502–1505 CrossRef CAS .
- Y. Li, H. Shao, Z. Lin, J. Lu, L. Liu, B. Duployer, P. O. Å. Persson, P. Eklund, L. Hultman, M. Li, K. Chen, X.-H. Zha, S. Du, P. Rozier, Z. Chai, E. Raymundo-Piñero, P.-L. Taberna, P. Simon and Q. Huang, Nat. Mater., 2020, 19, 894–899 CrossRef CAS .
- L. Liu, M. Orbay, S. Luo, S. Duluard, H. Shao, J. Harmel, P. Rozier, P.-L. Taberna and P. Simon, ACS Nano, 2022, 16, 111–118 CrossRef CAS .
- D. Gandla, Z. Zhuang, V. V. Jadhav and D. Q. Tan, Energy Storage Mater., 2023, 63, 102977 CrossRef .
- H. Ang, H. Wang, B. Li, Y. Zong, X. Wang and Q. Yan, Small, 2016, 12, 2859–2865 CrossRef CAS .
- H. Zhou, Z. Chen, E. Kountoupi, A. Tsoukalou, P. M. Abdala, P. Florian, A. Fedorov and C. R. Müller, Nat. Commun., 2021, 12, 5510 CrossRef CAS PubMed .
- R. B. Levy and M. Boudart, Science, 1973, 181, 547–549 CrossRef CAS PubMed .
- J. Halim, S. Kota, M. R. Lukatskaya, M. Naguib, M.-Q. Zhao, E. J. Moon, J. Pitock, J. Nanda, S. J. May, Y. Gogotsi and M. W. Barsoum, Adv. Funct. Mater., 2016, 26, 3118–3127 CrossRef CAS .
- Y. Guo, S. Jin, L. Wang, P. He, Q. Hu, L.-Z. Fan and A. Zhou, Ceram. Int., 2020, 46, 19550–19556 CrossRef CAS .
- A. Byeon, C. B. Hatter, J. H. Park, C. W. Ahn, Y. Gogotsi and J. W. Lee, Electrochim. Acta, 2017, 258, 979–987 CrossRef CAS .
- Y. Wang, Y. Nian, A. N. Biswas, W. Li, Y. Han and J. G. Chen, Adv. Energy Mater., 2021, 11, 2002967 CrossRef CAS .
- H. Fashandi, M. Dahlqvist, J. Lu, J. Palisaitis, S. I. Simak, I. A. Abrikosov, J. Rosen, L. Hultman, M. Andersson, A. Lloyd Spetz and P. Eklund, Nat. Mater., 2017, 16, 814–818 CrossRef CAS PubMed .
- M. Nechiche, T. Cabioc'h, E. N. Caspi, O. Rivin, A. Hoser, V. Gauthier-Brunet, P. Chartier and S. Dubois, Inorg. Chem., 2017, 56, 14388–14395 CrossRef CAS PubMed .
- W. Jiang, H. Shi, M. Shen, R. Tang, Z. Tang and J.-Q. Wang, ACS Appl. Mater. Interfaces, 2022, 14, 14482–14491 CrossRef CAS PubMed .
- Y. Hu, J. Zhou, L. Li, Z. Hu, T. Yuan, C. Jing, R. Liu, S. Xi, H. Jiang and J.-Q. Wang, J. Mater. Chem. A, 2022, 10, 602–610 RSC .
- H. Jo, Y. Yang, A. Seong, D. Jeong, J. Kim, S. H. Joo, Y. J. Kim, L. Zhang, Z. Liu and J.-Q. Wang, J. Mater. Chem. A, 2022, 10, 2271–2279 RSC .
- T. Shinagawa, A. T. Garcia-Esparza and K. Takanabe, Sci. Rep., 2015, 5, 13801 CrossRef PubMed .
- Z. W. Seh, K. D. Fredrickson, B. Anasori, J. Kibsgaard, A. L. Strickler, M. R. Lukatskaya, Y. Gogotsi, T. F. Jaramillo and A. Vojvodic, ACS Energy Lett., 2016, 1, 589–594 CrossRef CAS .
- G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169 CrossRef CAS PubMed .
- J. Hafner and J. Compu, Chem, 2008, 29, 2044–2078 CAS .
- J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS .
- G. Henkelman, B. P. Uberuaga and H. Jónsson, J. Chem. Phys., 2000, 113, 9901–9904 CrossRef CAS .
- M. Samanian and M. H. Ghatee, J. Mol. Model., 2021, 27, 268 CrossRef CAS .
- J. Kaczkowski, Acta Phys. Pol., A, 2012, 121, 1142–1144 CrossRef CAS .
- Y. Xie and J. A. Blackman, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 69, 172407 CrossRef .
- L. Dai, F. Yao, L. Yu, C. Fang, J. Li, L. Xue, S. Zhang, P. Xiong, Y. Fu and J. Sun, Adv. Energy Mater., 2022, 12, 2200974 CrossRef CAS .
- S. Jiao, X. Fu and H. Huang, Adv. Funct. Mater., 2022, 32, 2107651 CrossRef CAS .
- B. Wei, Z. Fu, D. Legut, T. C. Germann, Q. Zhang, S. Du, H. Zhang, J. S. Francisco and R. Zhang, J. Phys. Chem. C, 2021, 125, 4477–4488 CrossRef CAS .
- J. Yuan, Y. Liu, T. Bo and W. Zhou, Int. J. Hydrogen Energy, 2020, 45, 2681–2688 CrossRef CAS .
- C. Wang, H. Shou, S. Chen, S. Wei, Y. Lin, P. Zhang, Z. Liu, K. Zhu, X. Guo, X. Wu, P. M. Ajayan and L. Song, Adv. Mater., 2021, 33, 2101015 CrossRef CAS .
- F. Scharmann, G. Cherkashinin, V. Breternitz, C. Knedlik, G. Hartung, T. Weber and J. A. Schaefer, Surf. Interface Anal., 2004, 36, 981–985 CrossRef CAS .
- L. Dai, F. Yao, L. Yu, C. Fang, J. Li, L. Xue, S. Zhang, P. Xiong, Y. Fu, J. Sun and J. Zhu, Adv. Energy Mater., 2022, 12, 2200974 CrossRef CAS .
- J. Jia, T. Xiong, L. Zhao, F. Wang, H. Liu, R. Hu, J. Zhou, W. Zhou and S. Chen, ACS Nano, 2017, 11, 12509–12518 CrossRef CAS PubMed .
- Y.-Y. Chen, Y. Zhang, W.-J. Jiang, X. Zhang, Z. Dai, L.-J. Wan and J.-S. Hu, ACS Nano, 2016, 10, 8851–8860 CrossRef CAS .
- U. Yorulmaz, A. Özden, N. K. Perkgöz, F. Ay and C. Sevik, Nanotechnology, 2016, 27, 335702 CrossRef PubMed .
- H. Kim, B. Anasori, Y. Gogotsi and H. N. Alshareef, Chem. Mater., 2017, 29, 6472–6479 CrossRef CAS .
- V. Kamysbayev, A. S. Filatov, H. Hu, X. Rui, F. Lagunas, D. Wang, R. F. Klie and D. V. Talapin, Science, 2020, 369, 979–983 CrossRef CAS PubMed .
- V. Natu and M. W. Barsoum, J. Phys. Chem. C, 2023, 127, 20197–20206 CrossRef CAS .
- M. Anayee, C. E. Shuck, M. Shekhirev, A. Goad, R. Wang and Y. Gogotsi, Chem. Mater., 2022, 34, 9589–9600 CrossRef CAS .
- S. Wei, P. Zhang, W. Xu, S. Chen, Y. Xia, Y. Cao, K. Zhu, Q. Cui, W. Wen, C. Wu, C. Wang and L. Song, J. Am. Chem. Soc., 2023, 145, 10681–10690 CrossRef CAS .
- H. Ding, Y. Li, M. Li, K. Chen, K. Liang, G. Chen, J. Lu, J. Palisaitis, P. O. Persson and P. Eklund, Science, 2023, 379, 1130–1135 CrossRef CAS PubMed .
- J. M. McEnaney, A. R. Singh, J. A. Schwalbe, J. Kibsgaard, J. C. Lin, M. Cargnello, T. F. Jaramillo and J. K. Nørskov, Energy Environ. Sci., 2017, 10, 1621–1630 RSC .
- F. Meng, J. Qin, X. Xiong, H. Zhang, M. Zhu and R. Hu, Cell Rep. Phys. Sci., 2023, 4, 101307 CrossRef CAS .
- Y. Tang, C. Yang, X. Xu, Y. Kang, J. Henzie, W. Que and Y. Yamauchi, Adv. Energy Mater., 2022, 12, 2103867 CrossRef CAS .
- Y. Tang, C. Yang, Y. Tian, Y. Luo, X. Yin and W. Que, Nanoscale Adv., 2020, 2, 1187–1194 RSC .
- S. Oswald, Encyclopedia of Analytical Chemistry: Applications, Theory and Instrumentation, 2006 Search PubMed .
- F. Ando, T. Gunji, T. Tanabe, I. Fukano, H. D. Abruna, J. Wu, T. Ohsaka and F. Matsumoto, ACS Catal., 2021, 11, 9317–9332 CrossRef CAS .
- J. Suntivich, K. J. May, H. A. Gasteiger, J. B. Goodenough and Y. Shao-Horn, Science, 2011, 334, 1383–1385 CrossRef CAS PubMed .
- V. Giulimondi, S. Mitchell and J. Pérez-Ramírez, ACS Catal., 2023, 13, 2981–2997 CrossRef CAS PubMed .
|
This journal is © The Royal Society of Chemistry 2024 |