Leilei
Yin
ab,
Yuyan
Liu
c,
Shuai
Zhang
b,
Yongkang
Huang
c,
Qiang
Wang
*a,
Jin-Cheng
Liu
*b,
Chao
Gu
b and
Yaping
Du
*b
aInner Mongolia Academy of Science and Technology, Hohhot, Inner Mongolia 010010, China. E-mail: wangqiang@imast.ac.cn
bTianjin Key Lab for Rare Earth Materials and Applications, Center for Rare Earth and Inorganic Functional Materials, School of Materials Science and Engineering, National Institute for Advanced Materials, Nankai University, Tianjin, 300350, China. E-mail: ypdu@nankai.edu.cn; liujincheng@nankai.edu.cn
cCollege of Chemistry, Nankai University, Tianjin, 300071, China
First published on 15th April 2025
Optimizing both mass transport and electronic structure of the active component is of interest to obtain electrocatalysts with superior oxygen evolution reaction (OER) performance. Here, we miniaturized the classical NiFe-layered double hydroxides (NiFe-LDHs) and integrated them into S/N co-doped hollow hierarchical porous carbon (SNHPC) loaded with rare earth La single atoms (La SAs) to obtain nanoreactors. The unique carbon framework induced uniform deposition of LDH nanodots and ensured adequate exposure during electrocatalysis. The advantages of the carbon carrier for the local electric field and interfacial OH− layer density in the catalytic process were confirmed by finite element simulations. The well-designed NiFe-LDH@La SNHPC exhibited satisfactory activity (overpotential of 251 mV at 10 mA cm−2) and stability in alkaline media, exceeding those of commercial RuO2. Impressively, a cathode catalyst combining NiFe-LDH@La SNHPC with Pt/C can be stabilized in rechargeable zinc–air batteries (ZABs) for more than 350 h. Theoretical calculations indicated that the introduction of La SAs modified the electronic structures of the NiFe-LDH nanodots, activated lattice oxygen activity, optimized the adsorption strength of the intermediates, and reduced rate-determining step energy barriers in OER. This study provides guidance for the preparation and design of sub-microreactors and information on the strong electron interaction effects induced by rare earth species.
New conceptsThis work introduces a carbon nanoreactor integrating miniaturized NiFe-layered double hydroxides (LDH) and rare-earth La single atoms, specifically designed for efficient electrocatalytic oxygen evolution reaction (OER). Unlike conventional rare-earth doping in LDH, the stepwise integration strategy avoids potential structural distortions in the host matrix caused by direct La doping while enabling effective electronic interactions between the two components. This concept highlights the intrinsic contributions of the unique electronic structure of rare-earth elements, thereby circumventing additional structural strain effects. Combining experimental and theoretical approaches, the study provides detailed insights into the localized coexistence of La single atoms and NiFe-LDH nanodots and their crucial role in reducing the energy barriers of OER. Additionally, the research validates the potential of the nanoreactor's abundant mesoporous and cavity structures in enhancing the mass transport of multiple reactants and products during OER. This new concept provides a promising strategy for the rational design of high-performance LDH catalysts using rare-earth elements and will further inspire the research and application of nano-framework materials in the field of green energy. |
Integration with carbon has been shown to positively impact LDHs. Porous carbon enriched with heteroatoms (N, O, etc.) was used to further improve their electrical conductivity and induce high exposure.11–13 Notably, OER occurring at the gas–liquid–solid three-phase interface involves diffusion and mass transfer processes of multiple reactants, intermediates, and products.14,15 Therefore, rational design of doped multilayer carbon nanostructures with porous shell layers and internal spaces is crucial. Achieving LDHs with anchored active components and exposed sites is beneficial for realizing electrolyte permeation and creating abundant mass diffusion pathways.16,17 This is a very promising strategy for maximizing the utilization of LDH sites and improving the catalytic performance.
Furthermore, doping additional guest metal species to induce more flexible electronic structures in LDHs to modulate local charge transfer is an effective way to enhance their intrinsic activity.18–20 The strong electronic interactions between the host metal ions in the LDH framework and the guest metal species have been shown to manipulate the d-band center of LDHs, resulting in near-optimal binding energies of oxygenated intermediates during OER processes.21 Among many alternative guest metal species, rare earth (RE) elements with varying degrees of 4f orbital electron filling have received much attention.22–25 The unique localized 4f orbitals of RE species offer more possibilities to modulate the electronic structure of the active component by overlapping and hybridizing with the orbitals.26,27 However, the introduction of RE species into the LDH lattice is prone to induce additional unavoidable structural strain effects, which are caused by the large ionic size of rare earths.28–30 The LDH-related structural changes may inevitably interfere with the intrinsic contribution of electronic effects. Therefore, the rational construction of well-defined and easily investigated RE-LDH hybridization systems remains challenging.
Taking inspiration from the considerations mentioned above, we stepwise integrated ultra-miniaturized NiFe-LDH nanodots and atomically dispersed RE sites (La-N4 motifs) into a S-doped hierarchical porous hollow carbon framework to construct a nanoreactor that can be used for efficient OER. The carbon support was obtained by simple one-step pyrolysis of zeolitic imidazolate framework-8 (ZIF-8) with melamine. Its hollow cavity, enriched mesopores, and highly open-space structure can facilitate the utilization of active sites and electrolyte mass transfer in OER processes. The effectiveness of the designed nanoreactor was demonstrated by finite element method (FEM) simulations. Furthermore, theoretical calculations revealed the significance of additional S sites in the substrate in inducing uniform deposition of NiFe-LDH nanodots and enhancing the overall charge transfer efficiency of the nanoreactor. More importantly, the stepwise integration strategy avoided the possible structural distortion caused by the direct doping of NiFe-LDH with La because of its ionic radius, while achieving the effective electronic interactions between the two. These rational designs will improve electrocatalytic performance of the nanoreactor for OER. The final sample, NiFe-LDH@La SNHPC, required only 251 mV overpotential to reach a current density of 10 mA cm−2 in 1.0 M KOH medium, exceeding the performances of commercial RuO2 catalysts and a series of comparison samples. In addition, its reliable electrochemical activity and long-term cyclability of more than 350 h were verified in rechargeable ZABs.
Further in-depth studies were carried out to characterize the porosity of two carbon matrices with different morphologies using N2 adsorption–desorption measurements. As shown in Fig. S3a and c (ESI†), NC displayed a typical type-I isotherm with sharply increased nitrogen uptake at relatively low relative pressure, indicating the dominant microporosity.33,34 In contrast, NHPC revealed a type-IV isotherm with a well-defined hysteresis loop at a relative pressure of 0.6 < P/P0 < 1.0 and an abrupt increase in adsorption volume at low relative pressures (P/P0 < 0.02), implying the simultaneous existence of micropores and mesopores.17 Further, pore size distribution curves revealed the hierarchical porous structure of NHPC (Fig. S3d, ESI†), with mesopores concentrated at 4.1 nm and 13.3 nm. NC is dominated by a microporous structure with only a few mesoporous structures (Fig. S3b, ESI†). Even though NHPC has a smaller Brunauer–Emmett–Teller specific surface area (679.9 m2 g−1) than NC (1147.9 m2 g−1), its unique accessible hierarchical porous structure both provided space for the subsequent anchoring and domain limitation of the NiFe-LDH nanodots and facilitates the exposure of more active sites in the subsequent electrocatalytic process.9,35 Moreover, it is worth noting that the large specific surface area and porous hollow structure of the nanoreactor may facilitate the rapid transport and removal of the gaseous products of the OER at the three-phase interface. On the one hand, the abundant hierarchical porous structure can serve as a channel for gas diffusion, helping O2 bubbles to diffuse to the electrode surface and reducing the aggregation of bubbles on the catalyst surface. On the other hand, the larger specific surface area also helps to disperse the bubbles on the catalyst surface, reducing the probability of bubble coalescence and making it easier for bubbles to be removed from the catalyst surface.36–38 Raman spectra of NHPC and NC samples were used to reveal the degree of defects and graphitization (Fig. S4, ESI†). Two main peaks are shown at 1351 cm−1 and 1585 cm−1, attributed to D-band (defective carbon) and G-band (graphitic sp2 carbon) features, respectively. The degree of defects was estimated from the integrated intensity (ID/IG) of the D- and G-bands.39–41 NHPC and NC exhibited close ID/IG values, implying the presence of abundant defective carbon in both, which provided effective anchoring sites for the adsorption of Ni2+/Fe3+ during the subsequent growth of NiFe-LDH nanodots.11 The relatively high ID/IG values of NHPC may be related to more carbon defects caused by corrosive gases from melamine pyrolysis.42 The specificity of the structure of the NHPC sample motivated us to use it as a framework for a nanoreactor in subsequent procedures.
In the next stage, the La SAs were immobilized on NHPC by simple La3+ adsorption, secondary pyrolysis, and acid leaching treatments. This is facilitated by the evaporation of low-boiling Zn species in ZIF-8 during the initial carbonation process, leaving abundant of N sites for subsequent anchoring of additional metals.43,44 Notably, inspired by theoretical calculations, we introduced additional heteroatoms, S, into the process. The density of states (DOS) of La SNHPC demonstrated the presence of more electronic states near the Fermi level, which facilitates the electron transfer during the reaction processes and enhances its catalytic activity (Fig. S5, ESI†).45–47 Detailed structural characterization implies a successful homogeneous loading of atomically dispersed La on the SNHPC skeleton, with additional metal aggregates not detected (Fig. S6–S10, ESI†). The presence of S species did not adversely affect the formation of La SAs.
Finally, NiFe-LDH nanodots were then grown in situ from solution on the La SNHPC to obtain the ultimate nanoreactor (NiFe-LDH@La SNHPC). Detailed preparation processes are described in the ESI† Experimental section. The samples containing NiFe-LDH nanodots exhibited distinct diffraction peaks located at 11°, 23°, 35°, and 60°, which are attributed to (003), (006), (012), and (110) planes of hydrotalcite-like NiFe-LDH (Fig. 1b and Fig. S11, ESI†), respectively.9 The TEM image and dark-field TEM image showed that the nanoreactor retained the inherent hollow morphology of the precursor and was observed with distinct bright spots that were never observed in NHPC and La SNHPC (Fig. 1c and d). These uniformly dispersed bright spots in the substrate can be referred to as NiFe-LDH nanodots, further validating the feasibility of our carefully designed synergistic strategy of hierarchical pore restriction and S-site-assisted deposition. Upon comparing the TEM images in Fig. S12 and S13 (ESI†), significant aggregation of LDH nanodots was observed in the sample without S. In contrast, the growth of NiFe-LDH nanodots on the S-containing carbon substrate (SNHPC) was widely and homogeneously dispersed compared to that on NHPC. The S-assisted deposition was further supported by theoretical calculations (Fig. S14, ESI†), which showed binding energies between NiFe-LDH and the substrates with and without S. The results indicated that introducing S leads to a stronger interaction between NiFe-LDH and the substrate, effectively inhibiting the aggregation of NiFe-LDH and yielding more dispersed and homogeneous products. The energy dispersive X-ray spectroscopy (EDS) mapping exhibited a uniform distribution of Ni, Fe, La, S, O, and N on the carbon framework, implying that both NiFe-LDH nanodots and La SAs are highly dispersed in the carbon-based nanoreactor (Fig. 1d). The microscopic coexistence of the two was further observed by aberration-corrected high angle annular dark field scanning transmission electron microscopy (AC-HAADF-STEM). Irregular nanoparticles of about 5–10 nm can be observed in Fig. 2a (outlined with green dashed lines). The lattice fringes with an interplanar spacing of 0.24 nm were ascribed to the (006) facet of NiFe-LDH (bottom of Fig. 2b). Additionally, abundant La single atoms (outlined with yellow dashed circles) were observed around the NiFe-LDH nanodots, which exhibit larger sizes and higher brightness (top of Fig. 2b). In addition to coexisting with each other in the plane depending on the carbon substrate (region 3 in Fig. 2a), NiFe-LDH nanodots and La SAs were found to be in direct spatial contact (region 4 in Fig. 2a). This may be attributed to the fact that the nucleation and growth of LDH nanodots on the carbon support is stochastic. The corresponding 3D fitting plot more visually illustrates the coexistence of the two previously mentioned states (Fig. 2c). The La sites with higher signal strength are marked with arrows in the figure. The localized effective coexistence state of the two species was further strongly evidenced in the EDS mapping images at higher magnification (Fig. 2d). Furthermore, Fig. S15 (ESI†) shows the microstructure of NiFe-LDH@SNHPC without La, where only the presence of NiFe-LDH nanodots was observed. The content of metals Ni, Fe, and La in NiFe-LDH@La SNHPC were determined to be 15.2 wt%, 4.7 wt%, and 0.5 wt%, respectively, by inductively coupled plasma-optical emission spectroscopy analysis (ICP-OES).
![]() | ||
Fig. 2 (a) AC-HAADF-STEM image of NiFe-LDH@La SNHPC. (b) and (c) Integrated pixel intensity profiles (b) and 3D fitting map (c) of different regions marked in Fig. 2a. (d) AC-HAADF-STEM image and the corresponding EDS elemental mapping of NiFe-LDH@La SNHPC at high magnification. (e) La L3-edge Fourier transformation of the EXAFS spectra in R space for NiFe-LDH@La SNHPC and La2O3. (f) La L3-edge EXAFS fitting curves of the NiFe-LDH@La SNHPC. Inset: Proposed architecture. The blue, gray, red, and yellow balls represent N, C, La, and S atoms, respectively. (g) and (h) High-resolution XPS spectra of Ni 2p (g) and Fe 2p (h) for different samples. |
To further affirm the local coordination environments of La sites in NiFe-LDH@La SNHPC catalysts, the X-ray absorption near-edge structure (XANES) and extended X-ray absorption fine structure (EXAFS) were exploited and analysed. The La L3-edge XANES spectra (Fig. S16, ESI†) verified that the valence of La single atoms was close to that of La2O3 (+3). Fig. 2e shows the Fourier transformation (FT)-EXAFS (R space) for the La L3-edge of NiFe-LDH@La SNHPC and La2O3 references. The La–N bonds were located at 2.1 Å and no La–La bond (3.7 Å) was observed in NiFe-LDH@La SNHPC, further verifying the atomic dispersion of La, consistent with the results of the AC-HAADF STEM. The wavelet transform EXAFS (WT-EXAFS) contour plots of NiFe-LDH@La SNHPC exhibited a major peak near 3.0 Å−1, unlike the La2O3, providing further evidence that La is atomically dispersed and coordinated to the light element N (Fig. S17, ESI†). According to the detailed fitting results (Table S1 and Fig. S18, ESI†), a single La atom was effectively anchored by four N atoms in NiFe-LDH@La SNHPC to form the La–N4 motifs (inset of Fig. 2f). There was no significant La–Ni/Fe bonding.
X-ray photoelectron spectroscopy (XPS) was used to further analyze the surface chemical state of the sample and possible electronic interactions. The chemical states of C, N, O, S, and La in NiFe-LDH@La SNHPC are explained in Fig. S19 (ESI†). The C 1s (Fig. S19a, ESI†) spectrum can be split into five configurations containing C–S (284.4 eV), sp2-C (284.8 eV), sp2-C (285.5 eV), C–N (286.6 eV), and C–O (288.8 eV), confirming the S doping.48 The N 1s spectrum (Fig. S19b, ESI†) can also be deconvoluted into four types of N species including pyridinic N (398.1 eV), pyrrolic N (399.3 eV), graphitic N (400.5 eV), and oxidized-N (402.2 eV).39 The O 1s spectrum (Fig. S19c, ESI†) shows that the strong peak at 531.1 eV in NiFe-LDH@La SNHPC represents the metal–hydroxide (M–OH) bond in NiFe-LDH.18 Furthermore, the S 2p spectra (Fig. S19d, ESI†) can be attributed to the S 2p3/2 (163.2 eV), S 2p1/2 (164.5 eV), and sulfate species (C–SOx).49 The lack of characteristic metal–S bond peaks implied that the foreign S atoms tend to bind to the carbon substrate rather than directly to the La atoms in the center, which is consistent with the EXAFS results. Notably, due to the lower content of La and the encapsulation effect of the carbon carrier, it was difficult to accurately identify the oxidation state of the La atoms using the attenuated photoelectron signal (Fig. S19e, ESI†). Focusing on the influence of the La sites on the electronic structure of Ni and Fe in LDH nanodots, the detailed fitting results indicated that Ni species were mainly present in the NiFe-LDH nanodots as Ni2+, while the trivalent cation positions were mainly contributed to by Fe species (Fig. 2g and h).9,50–52 The Ni 2p and Fe 2p peaks in NiFe-LDH@La SNHPC were positively shifted by 0.3 and 0.5 eV, respectively, compared to NiFe-LDH@SNHPC, suggesting that electrons were transferred from NiFe-LDH nanodots to La SAs, by a notable amount. This suggested that based on the two coexisting states of La SAs and NiFe-LDH nanodots mentioned earlier, the charge transfer between the two contributes to increasing the oxidation state of Ni and Fe, thus enhancing their intrinsic activity towards OER.18,53 The calculated charge density difference plots of the NiFe-LDH@La SNHPC heterostructure (Fig. S20, ESI†) show that the charge accumulates on the La SNHPC side and is depleted on the NiFe-LDH side, which indicated spontaneous electron migration from the NiFe-LDH side to the substrate. Specifically, based on the Bader charge analysis, NiFe lost 0.013 e− per atom after the introduction of La sites and results in the increase of oxidation states, which agrees well with the XPS analysis.
Further finite element analysis, based on the morphology and dimensions of the catalyst, showed that NiFe-LDH@La SNHPC exhibited significantly better localized OH− enrichment and electrostatic field strength enhancement in the catalyst microenvironment than NiFe-LDH@La SNC with a single microporous solid structure. This improvement can be attributed to its unique cavity and hierarchical pore structure (Fig. 3d and e). The higher electrostatic field strength promoted the adsorption of water molecules on the catalyst surface and further enhanced the local electron transport, which directly benefited the OER kinetics.54,55 In addition, the accumulated OH− can effectively alleviate the delayed replenishment of the heavily consumed OH− reactants during the OER process.56–58 It is worth emphasizing that the gas produced after the OER can be transported through the channels inside the nanoreactor and the intermediate chamber, and the mass transfer process can be further accelerated.
The double-layer capacitance (Cdl) calculated from cyclic voltammetry (CV) curves at different scan rates in the non-Faraday current region was positively correlated with the electrochemical active surface area (ECSA), which was an important indicator of the actual number of active centers exposed to the electrocatalyst and represented the inherent activity level of the catalyst (Fig. S24, ESI†).59–61 The Cdl value for NiFe-LDH@La SNHPC is 3.25 mF cm−2, which is higher than those for NiFe-LDH@La NHPC (2.53 mF cm−2) and NiFe-LDH@NHPC (1.38 mF cm−2), implying that NiFe-LDH@La SNHPC has a larger ECSA. A larger ECSA was thought to be favorable for providing an abundance of more exposed active sites in the OER and higher intrinsic activity.28 The charge transfer process of the electrode reaction was further evaluated using electrochemical impedance spectroscopy (EIS). The radius of curvature was directly proportional to the charge transfer resistance and inversely proportional to the charge transfer capacity in the low-frequency portion of Fig. S25 (ESI†). The NiFe-LDH@La SNHPC catalyst also exhibited the lowest values in these samples, supporting its efficient charge transfer capacity.55 The durability of the catalyst was also evaluated. After running the chronoamperometric i–t measurement at a current density of 50 mA cm−2 for up to 30 h, NiFe-LDH@La SNHPC still retained 91.1% of the initial current, which was significantly better than RuO2 (Fig. 3g). It also shows only 25.6% performance loss at a higher current density 100 mA cm−2 (Fig. S26, ESI†). The LSV curves before and after the 2000 CV cycles in Fig. S27 (ESI†) also showed only slight changes from each other, indicating excellent OER durability in alkaline solutions. In addition, its original morphological characteristics and uniform elemental distribution were maintained after the stability test, and no obvious structural collapse and active component aggregation were observed (Fig. S28 and S29, ESI†). XPS spectra were recorded after 2000 CV cycles to assess the elemental electronic states (Fig. S30, ESI†). Ni 2p high-resolution spectra exhibited characteristic Ni3+ peaks that differed from those of the pristine catalysts, indicating that the surface metal Ni species had been rearranged and oxidized during the OER process. The peak positions of the Fe 2p XPS spectra showed little shift after cycling. In addition, the fact that the changes in the elements contained in the carbon framework could not be observed implied overall structural stability during the catalytic process.
Inspired by the excellent OER performance of NiFe-LDH@La SNHPC, a two-electrode electrolyzer is assembled using commercial 20% Pt/C and the NiFe-LDH@La SNHPC sample as the cathode and anode, respectively, in 1.0 M KOH for the overall water splitting. The NiFe-LDH@La SNHPC‖Pt/C couple achieved a prominent water splitting performance with low cell voltages of 1.55 and 1.73 V at 10 and 50 mA cm−2, respectively, lower than that of RuO2‖Pt/C (1.58 V@10 mA cm−2, 1.84 V@50 mA cm−2) (Fig. S31a, ESI†). Multi-step chronopotentiometry tests showed that NiFe-LDH@La SNHPC at different current densities exhibits lower voltages than commercial RuO2. In contrast to the obvious current density decay of RuO2, the change for NiFe-LDH@La SNHPC was minimal (Fig. S31b, ESI†). Furthermore, its good durability was also reflected in the stable operation for up to 30 h at a current density of 50 mA cm−2 (Fig. S31c, ESI†).
The performance of NiFe-LDH@La SNHPC in a rechargeable ZAB was also evaluated. Prior to this, the oxygen reduction reaction (ORR) performance of NiFe-LDH@La SNHPC in 0.1 M KOH was investigated and commercial 20% Pt/C was used as a comparison sample. The LSV curves showed that NiFe-LDH@La SNHPC possessed a half-wave potential (E1/2) of 0.76 V (Fig. S32a, ESI†), lower than that of Pt/C (0.85 V). The Koutecky–Levich (K–L) plots obtained from the LSV curves of the sample at different rotation rates showed that the ORR reaction proceeded by a four-electron process (Fig. S32b, ESI†). The ORR stability was also subsequently assessed (Fig. S32c, ESI†).
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5mh00313j |
This journal is © The Royal Society of Chemistry 2025 |