Peng
Liu
a,
Jiahui
Ye
a,
Kuan
Deng
a,
Xuesong
Liu
a,
Haohui
Dong
a,
He
Zhang
a,
Wen
Tian
a and
Junyi
Ji
*ab
aSchool of Chemical Engineering, Sichuan University, Chengdu 610065, P. R. China. E-mail: junyiji@scu.edu.cn
bState Key Laboratory of Polymer Materials Engineering, Sichuan University, Chengdu 610065, P. R. China
First published on 23rd January 2025
Single-atom catalysts (SACs) have attracted considerable interest in the field of electrocatalysis due to their high efficiency of metal utilization and catalytic activity. However, traditional methods of SACs fabrication are often complex and time-consuming. Herein, F–Ru@TiOxNy was synthesized using a straightforward and universal approach via in situ surface etching and heteroatoms immobilization on a vacancies-rich hierarchical TiOxNy nanorods array. The fluorine ion-etched TiOxNy nanorods could produce abundant oxygen vacancies and F–Ti/F–C bonds, which could capture and stabilize Ru heteroatoms by strong host–guest electronic interactions. Due to the synergistic effect of oxygen vacancies anchoring and F–C bonds-assisted stabilization of single atoms, F–Ru@TiOxNy revealed excellent electrocatalytic hydrogen evolution performance, a low overpotential of 20.8 mV at 10 mA cm−2, a Tafel slope of 59.9 mV dec−1 and robust stability at 100 mA cm−2 over 48 h. Furthermore, this universal strategy could be applicable to various heterometals (Pd, Ir, Pt), which also exhibited high heteroatoms dispersity and high electrocatalytic HER activity/stability. This fabrication method is simple, easy-scalable and versatile, showcasing significant potential for electrocatalysts design and promising application prospects in electrocatalytic energy conversion.
Highly dispersed precious metal catalysts demonstrate exceptional electrocatalytic activity and stability, which significantly diminishes the required quantity of precious metals. Especially, single atom catalysts (SACs)5 are extensively employed in electrocatalytic hydrogen evolution due to their high activity, favourable selectivity, and ultralow amount of precious metals consumed. However, traditional fabrication methods for SACs are usually complex and time-consuming, with weak interfacial force between heteroatoms and substrate, leading to an aggregation trend of precious metals and limitation of large-scale fabrication.6–8 Among various strategies for SACs fabrication, the defect engineering strategy is widely applied.9 It involves constructing defect sites on the surface of a support (e.g., carbon defects,10,11 oxygen defects,12 sulfur defects,13,14 and metal defects),15 then utilizing the support defects as “traps” to capture and separate the individual heteroatoms. Subsequently, the well dispersed single atoms can be stabilized via electronic modulation between the heterometal atoms and defect sites due to the different electron energy levels. This approach achieves uniform dispersion and local stabilization of heteroatoms to prevent aggregation.16 Therefore, it is crucial to effectively create and utilize defects on the substrate to capture and stabilize the heteroatoms, and further regulate their coordination environment to improve the electrocatalytic activity, selectivity, and stability.17–20 Thus, various approaches have been developed to optimize the electronic structure of Ru, such as introducing a cation/anion,21 designing a specific structure,22 and regulating the support structure,23 to achieve excellent electrocatalytic performances. However, little attention has been paid to the halogen effects on the ligands engineering. Fluoride ions (F−) possess the highest electronegativity and a highly malleable electronic structure, which can optimize the adsorption energy of critical key intermediates through enhanced electronic interactions. Meanwhile, the strong electronegativity of fluorine atoms can enhance the stability of anchored heteroatoms by effectively adsorbing positive-charged metal ions, thus further strengthening heteroatoms–support interactions.24–26 Moreover, the surface wettability of the electrode can also be modulated by adjusting the surface energy, thus facilitating the infiltration of the electrolyte and removal of the produced H2 gas.4,19,22 Therefore, it is meaningful and feasible to attempt oxygen vacancy anchoring and F− doping strategies to prepare highly stable SACs for electrocatalytic hydrogen evolution.
Herein, we presented a straightforward and universally applicable method for the preparation of heterometal single-atom catalysts on a free-standing hierarchical TiN/TiO2 heterostructure nanorods array (F–M@TiOxNy, M = Ru, Pd, Ir, Pt). The F–M@TiOxNy catalysts were synthesized using a simple and easy-scalable in situ etching and immobilization technique. Fluorine ion-etched TiOxNy nanorods with abundant oxygen vacancies and F–C/F–Ti bonds were utilized to capture and stabilize the heteroatoms. Detailed characterization indicated heteroatoms to be anchored within the oxygen vacancies in TiOxNy and F–C bonds could stabilize the heteroatoms, while the surface immobilization on hierarchical TiOxNy nanorods could effectively expose active sites and reduce the precious metal dosage. Therefore, due to the synergistic effect of oxygen vacancy anchoring and F–C assisted stabilization, F–Ru@TiOxNy revealed high electrocatalytic performance for hydrogen evolution with a low overpotential of 20.8 mV at 10 mA cm−2, a Tafel slope of 59.9 mV dec−1 and a robust stability at 100 mA cm−2 over 48 h. Furthermore, this universal method is applicable to various heterometals (Pd, Ir, Pt), which also exhibit good heteroatoms dispersity and high electrocatalytic HER activity/stability. Therefore, this simple and easy-scalable fabrication method showcases significant potential and promising application prospects due to its mild synthetic conditions and versatility.
The experimental details pertaining to the characterization of materials and the electrochemical testing of these composites are provided in the ESI experimental section in the ESI.†
![]() | ||
Fig. 1 An in situ fluoride ion etching and heteroatoms immobilization strategy for the synthesis of F–M@TiOxNy (M = Ru, Pd, Ir, Pt) (schematic). |
Scanning electron microscopy (SEM) was employed to examine the morphology evolution of the composites. Initially, TiO2 nanorod arrays with a smooth surface anchored onto the carbon fiber surface were synthesized using a seed-assisted hydrothermal method (Fig. 2a). To induce oxygen vacancy defects and a carbon coating layer onto the nanorods surface, TiOxNy was produced by nitridation of TiO2 through pyrolysis of DCDA under a high temperature, which resulted in the rough surface and hollow structure of the nanorods (Fig. 2b). During solution immersion, the morphology of the nanorod arrays was well-preserved (Fig. 2c). The rough surface with a carbon coating layer enhanced the number of immobilization points for heteroatoms and increased the number of accessible active sites for the HER. Importantly, significant heterometal agglomeration was not observed on the TiOxNy nanorods surface, demonstrating the uniform distribution of the heteroatoms.
Transmission electron microscopy (TEM) was employed to evaluate the microstructure of the composites. As illustrated in Fig. 2d, TiOxNy nanorods exhibited a porous and hollow structure, while the rough surface was consistent with the SEM image. The lattice spacing of 0.211 nm corresponded to the (200) crystal plane of TiN (Fig. 2d), demonstrating the successful nitridation of the nanorods. The in situ phase transformation could lead to the formation of abundant crystalline dislocations and defects to facilitate subsequent heteroatoms immobilization. After surface etching and heteroatoms immobilization, the microstructure of the TiOxNy nanorods remained almost unchanged (Fig. 2e and f), indicating the mild surface etching process and robust structure stability of the nanorods. Furthermore, the immobilization and uniform dispersion of Ru single atoms was corroborated by the red spots observed in the high-resolution image. The high-resolution TEM (HRTEM) image of F–Ru@TiOxNy nanoparticles revealed a bright and ordered arrangement of Ti atoms along the (111) crystal direction (Fig. 2g). Both the three-dimensional pseudo-colour surface intensity plot and X–Y intensity profiles of the F–Ru@TiOxNy lattice exhibited the atomically dispersed feature of Ru element. Furthermore, the additional bright atomic columns on the TiOxNy surface could be identified as Ru atoms due to the lower atomic number of Ti compared with that of Ru.30 The introduction of Ru heteroatoms could introduce new active sites and consequently enhance the surface activity of the electrocatalyst for the HER.31 Additionally, the energy-dispersive X-ray (EDX) elemental line scan and mapping images indicated the homogeneous distribution of Ti, O, N, C, F, and Ru elements without obvious discernible aggregation of Ru species (Fig. 2h and S14†), indicating the successful introduction and stable immobilization of Ru and F species. Therefore, these results demonstrated the synthesis of the TiOxNy composite with Ru single-atoms loading via immersion in a precursor solution.
The crystallinity and phase structures of the as-prepared composites were analyzed by X-ray diffraction (XRD). As depicted in Fig. 3a, the nitridation process resulted in a weakening of the TiO2 peaks and the presence of new peaks associated with TiN, while the crystalline phases of TiO2 and TiN corresponded closely with rutile TiO2 (JCPDS 77-0440) and osbornite TiN (JCPDS 38-1420), respectively. In contrast, for F–Ru@TiOxNy, the peak intensity remained almost unchanged and no new peaks could be observed, suggesting weak surface damage during the etching process and ultralow amount of Ru loading. Furthermore, as shown in electron paramagnetic resonance (EPR) curves, the TiOxNy exhibited an obvious higher oxygen vacancies intensity compared with that of TiO2 (Fig. 3b), which was attributed to the nitridation process that induces many crystalline interfaces, dislocations and vacancies. It is worth noting that fluoride ion etching led to further enhancement of the oxygen vacancies strength in F–TiOxNy (Fig. S15a†), demonstrating the strong surface reaction/interaction during surface etching and the introduction of extra vacancies.32 Meanwhile, the oxygen vacancies intensity revealed a significant decrease with the presence of Ru ions (Fig. S15b†), which may have been due to the anchoring of Ru heteroatoms into oxygen vacancies. Moreover, as the concentration of Ru salt in the etching solution increased, the peak intensity of oxygen vacancies decreased accordingly (Fig. S15c†), demonstrating the accurate location and strong interaction of the Ru heteroatoms with surface vacancies.33
Moreover, the oxidation valence states of the surface elements were investigated by X-ray photoelectron spectroscopy (XPS). The full XPS survey revealed that F–Ru@TiOxNy displayed two additional peaks compared with TiOxNy, indicating the presence of F and Ru elements (Fig. S16†). The F 1s spectra of F–Ru@TiOxNy could be deconvoluted into three peaks located at 687.78, 684.78 and 683.78 eV (Fig. 3c), which corresponded to semi-ionic F–C,25 F–Ti,34 and covalent F–C,35 respectively. In comparison, F–Ru@TiO2 revealed only covalent F–C, demonstrating the high surface reactivity and strong interaction of F ions with the defect-rich Ti sites and amorphous carbon. It is worth noting that the formation of semi-ionic C–F bonds tends to adsorb H2O molecules onto the hydrophilic end F–C through hydrogen bonding, so the introduction of F enhances the wettability of the electrolyte on the electrode surface.25 The presence of F–Ti bonds is beneficial for regulating the electronic structure and coordination environment of Ru heteroatoms,25 which provides an extra interaction force to guarantee stability during electrocatalytic hydrogen evolution. Furthermore, O 1s in F–Ru@TiOxNy could be deconvoluted into four peaks centred at binding energies of 533.03, 532.08, 531.38 and 530.73 eV (Fig. 3d), which were composed to C–O, adsorbed oxygen (OA), oxygen vacancy (Ov) and lattice oxygen (OL), respectively.36–39 The O 1s spectra shift towards higher binding energy after F− doping, so the electronic environment around O/N is highly weakened due to the high electronegativity of F atoms.40 Meanwhile, the strong positive peak shift in O 1s may also due to the significant electron transfer from Ti to Ru heteroatoms.41 Moreover, the proportion of the Ov peak area significantly decreased after immobilization of Ru, indicating that Ru atoms were anchored within the oxygen vacancies to avoid aggregation, which is in line with the EPR results. As illustrated in Fig. 3e, the N 1s spectra of TiOxNy and F–Ru@TiOxNy consisted of similar peaks located at 402.78, 399.88, 397.38 and 396.33 eV, which corresponded to N–C, N–O, O–N–Ti, and N–Ti, respectively. Obviously, the peak intensities of O–N–Ti and N–Ti in F–Ru@TiOxNy revealed a significantly decreased trend, possibly due to the formation of F–Ti bonds.
As shown in the Ti 2p and Ru 3p spectra in Fig. 3f, the Ru 3p peak was obviously weak without surface F− etching, which demonstrated the necessity of the presence of F species and extra defects to anchor sufficient Ru atoms. In comparison, with the addition of F− etching, a small number of Ru atoms could be introduced into TiO2 lattices. After F− etching of TiOxNy, the loading amount of Ru heteroatoms was greatly increased, while an obvious shift in Ti 2p orbitals and a significant decrease in peak intensity could be observed, indicating the doping of fluoride ions. The Ti 2p and Ru 3p spectra in F–Ru@TiOxNy exhibited 10 prominent peaks located at 485.74/463.21, 484.19/461.91, 463.51/457.07, 462.55/455.64, 461.76/454.29 eV, which aligned well with Ru 3p1/2, Ru 3p3/2, Ti–O, Ti–N–O and Ti–N,42,43 respectively. Moreover, the C 1s and Ru 3d orbitals also revealed similar situations; no F–C bond could be observed in Ru@TiO2 without the addition of the KF etchant (Fig. 3g). With the presence of KF, F–C bonds could be clearly observed on F–Ru@TiO2, while the carbon species may have originated from carbon fiber. In comparison, the peak intensity of F–C bonds and Ru 3d peaks of F–Ru@TiOxNy were significantly enhanced, while the peaks around 295.46 and 292.68 eV demonstrated the formation of CF2 and CF, respectively.44 The formation of F–C bonds and many oxygen vacancies during KF etching could enhance the host–guest interaction force to immobilize Ru heteroatoms, thereby increasing the loading amount and stability of single atoms. The binding energies located at 296.04, 293.28, 287.91 and 284.80 eV were associated with CO, C–O, C–N and C–C, respectively. Meanwhile, the peaks centred at 286.80/283.13 and 285.70/282.08 eV were attributed to Run+ and Ru0,23 respectively. Both the Run+ and Ru0 peaks shifted to lower binding energies, suggesting the enhanced electron-acquiring behaviour of Ru sites.25 Moreover, the incorporated fluorine could also enhance the surface hydrophilicity of the electrocatalyst, significantly improving the bubble diffusion kinetics at the catalyst surface and thus boosting the catalytic stability (Fig. S17†). Therefore, the formation mechanism of the SACs by etching TiOxNy hollow nanorods with fluoride ions could be attributed to the immobilization of Ru atoms within the oxygen vacancy defects, as well as the electronic structure regulation of Ru atoms by Ru–Ti interactions and F–C/F–Ti bonds.
Moreover, operando electrochemical impedance spectroscopy (EIS) at various voltages was conducted to elucidate the surface adsorption and desorption kinetics of reaction intermediates on Ru active sites. The fitting results of relevant EIS are shown in Fig. 4d–f and S19.† Generally, the equivalent impedance gradually decreases with an increased applied potential. The Nyquist plots of F–Ru@TiOxNy revealed the smallest semicircular trend than that of Pt/C and F–Ru@TiO2 under high applied potentials (Fig. 4d–f). This suggested that the F–Ru@TiOxNy surface was favorable for the adsorption of reactants and intermediates during the HER, which was attributed to the optimized low-valence state Ru sites with electronic structures matching the reaction intermediates.49 Compared with Pt/C, F–Ru@TiO2 also exhibited a smaller semicircle, indicating that the slight surface etching of TiO2 by fluoride ions could also load a small number of Ru sites and favor the surface adsorption/reaction. In addition, F–Ru@TiOxNy revealed the smallest surface transfer resistance (Rs) and charge transfer resistance (Rct) under a high potential, which is beneficial for surface adsorption and the electrocatalytic HER. As shown in the Bode plots in Fig. 4g–i, the response of phase transition peaks was derived from the interfacial charge transfer and inner-layer electron transport of the electrocatalyst, which is consistent with the Volmer step and Heyrovsky reaction, respectively.50 The first step of the HER is the Volmer reaction with H2O in an alkaline medium to generate H*. The second step is the rate-determining step (RDS), which depends on the activity of the catalyst. For Ru-based catalysts, the second step is the Heyrovsky reaction, where H* reacts with water to produce H2.51 Compared with Pt/C and F–Ru@TiO2, in situ EIS indicated a faster decrease trend of the phase angle for F–Ru@TiOxNy, indicating the faster Volmer step and Heyrovsky step. Therefore, the Ru single atom can accelerate the hydrolysis dissociation rate and rapidly generate H2 through the Heyrovsky step.52 Moreover, the rapid decrease of the phase angle peak of F–Ru@TiOxNy in the low/high-frequency region verified the facilitated charge transfer efficiency at the electrolyte/catalyst interface and the rapid transfer of electrons within the bulk body, which could effectively stimulate the HER reaction kinetics. Moreover, the equivalent circuit models of F–Ru@TiOxNy for the operando EIS measured in 1 M KOH are shown in Fig. 4j. Generally, the equivalent resistance of these electrocatalysts gradually decreased with an increment in applied potential, indicating the accelerated charge transfer efficiency with the increasing potential. Below −0.944 V, the equivalent impedance of all catalysts was relatively similar, but that of the F–Ru@TiOxNy decreased rapidly to the lowest values above −0.964 V, indicating the rapidly decreased charge/electron transfer resistance of F–Ru@TiOxNy. After achieving −1.104 V, the equivalent resistance of R2 for F–Ru@TiOxNy reflecting the HER reaction dropped to near zero, indicating the optimized surface adsorption and reaction kinetics towards hydrogen production.53 Furthermore, F–Ru@TiOxNy demonstrated high long-term stability due to the strong interface electronic interaction, and a negligible decrease in current density around 10 mA cm−2 at constant applied potential for over 24 h could be achieved (Fig. 4k), which outperformed F–Ru@TiO2 and Pt/C with a relatively high attenuation trend. Even under a high current density of ∼100 mA cm−2, F–Ru@TiOxNy revealed excellent stability over 48 h, while the LSV curves before/after 48 h exhibited almost complete overlap (Fig. 4k inset), further demonstrating the robust and stable electrocatalytic HER activity. Moreover, F–Ru@TiOxNy retained its nanorod array structure after stability testing, but the surface underwent reconstruction under long-term reductive potential, resulting in the formation of nanoneedles or nanosheets arrays (Fig. S20†).54 Furthermore, no significant change could be observed for crystalline phases after the i–t test (Fig. S21a†). However, the peak intensity of N 1s, F 1s and C 1s and Ru 3d decreased obviously, while the O 1s peak showed a significant shift and the Ti 2p peak tended towards a non-nitridation state (Fig. S21b–f†), demonstrating that the surface reconstruction structure may have originated from amorphous TiOx. Therefore, F–Ru@TiOxNy revealed high HER activity and Ru heteroatoms immobilization stability during the long-term test.
To further illustrate the electrocatalytic performance of F–M@TiOxNy (M = Pd, Ir, and Pt) synthesized via this universal synthetic method, the electrocatalytic HER activity was evaluated in 1 M KOH solution. As depicted in Fig. 6a, all these F–M@TiOxNy composites demonstrated remarkable electrocatalytic activity for hydrogen production. Specifically, F–Pd@TiOxNy, F– Ir@TiOxNy, and F–Pt@TiOxNy exhibited low overpotential of 18.9, 13.2, and 22.3 mV at 10 mA cm−2, and 47.5, 53.9, and 59.5 mV at 100 mA cm−2, respectively. Moreover, the ECSA for F–Pd@TiOxNy, F–Ir@TiOxNy, and F–Pt@TiOxNy was calculated to be 248.5, 127.6, and 214.5 mF cm−2, respectively (Fig. 6b and S24a–c†), which was significantly higher than that of TiOxNy (98.4 mF cm−2). The underpotentially deposited hydrogen (Hupd) desorption peaks of F–M@TiOxNy (M = Pd, Ir, and Pt) moved to negative potential (Fig. S24d†), verifying a favourable hydrogen desorption step due to the relatively weak chemisorption strength of H*. Furthermore, the F–M@TiOxNy (M = Pd, Ir, Pt) composites exhibited relatively low charge transfer impedance (Fig. 6c), which was obviously lower than that of commercial Pt/C (Fig. 4d). The Tafel slopes for F–Pd@TiOxNy, F–Ir@TiOxNy, and F–Pt@TiOxNy were measured to be 41.5, 48.6, and 46.1 mV dec−1, respectively (Fig. 6d), together with the favourable hydrogen desorption kinetics (Fig. S24d†), demonstrating fast surface adsorption/desorption efficiency and reaction kinetics. As shown in Fig. S25,† the F–M@TiOxNy electrocatalysts revealed an obviously larger enclosed area for CV curves compared with the other contrast composites, demonstrating a distinct difference in accessible surface area. Meanwhile, the turnover frequency (TOF) of F–M@TiOxNy (M = Ru, Pd, Ir, Pt) was close to or better than that of commercial Pt/C (Fig. 6e). The potentials required for F–Ir@TiOxNy, F–Pd@TiOxNy, F–Pt@TiOxNy, F–Ru@TiOxNy and Pt/C to achieve a TOF of 1.5 s−1 were 86.98, 121.37, 209.48, 151.88 and 164.36 mV, respectively, which illustrated the high intrinsic electrocatalytic activity of F–M@TiOxNy. Furthermore, the long-term stability of F–M@TiOxNy composites was also evaluated at a constant applied potential of ∼100 mA cm−2, and relatively high long-term stability could be maintained over 24 h (Fig. 6f). The LSV curves before and after 24 h revealed almost complete overlap (Fig. S26†), indicating stable reaction kinetics and heteroatoms interface immobilization stability. Furthermore, compared with the recently reported high-end Ru-based HER electrocatalysts, the F–M@TiOxNy (Ru, Pd, Ir, and Pt) composites demonstrated significantly lower onset overpotentials and Tafel slopes, and excellent stability (Fig. 6g and Table S1†), demonstrating superior HER activity and stability, as well as fast reaction kinetics due to the high dispersity and stability of the heteroatoms.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4sc08603a |
This journal is © The Royal Society of Chemistry 2025 |