Fang-Yi Lia,
Shan Guanb,
Jianming Liua,
Changhao Liua,
Junfeng Zhangb,
Ju Guc,
Zhaosheng Li
a,
Zhigang Zoua and
Zhen-Tao Yu
*a
aNational Laboratory of Solid State Microstructures and Jiangsu Provincial Key Laboratory for Nanotechnology, College of Engineering and Applied Sciences, Nanjing University, Nanjing, Jiangsu 210093, China. E-mail: yuzt@nju.edu.cn
bState Key Laboratory of Engines, School of Mechanical Engineering, Tianjin University, Tianjin, 300350, China
cSchool of Physics, Nanjing University, Nanjing, 210093, People's Republic of China
First published on 23rd June 2025
The development of low-cost transition metal catalysts for use in alkaline water electrolysis (AWE) at high current densities is essential for achieving high-performance water splitting. Here, we reported a CrSb–MnO2 catalyst, which shows a low overpotential of 263 mV at 100 mA cm−2 and outstanding stability with only a small degradation of the catalyst after 100h of operation at 1 A cm−2 (1 M KOH). In addition, the catalyst also achieved excellent performance in AWE (1.69 V@1 A cm−2). This enhanced performance is not only due to lattice-strain engineering, which effectively modulates the electronic configurations of the active sites, but also due to bimetallic synergy, which improves the dynamics of metal–metal charge transfer. In situ differential electrochemical mass spectrometry (DEMS) and Fourier-transform infrared (FTIR) analyses revealed that the CrSb–MnO2 catalyst preferred the adsorbate evolution mechanism (AEM) during the alkaline OER. This preference contributes to sustained stability under high current conditions in alkaline media. This work offers a novel approach for designing membrane electrodes that can operate efficiently and stably under large currents.
Broader contextElectrochemical water splitting is an advanced technology for the large-scale production of renewable green hydrogen. However, the oxygen evolution reaction (OER) at the anode involves a slow four-electron transport process, which significantly restricts the overall oxygen evolution rate and the efficiency of water oxidation. The MnO2 catalyst is widely utilized in alkaline electrolyzers due to its abundant valence states and low cost. Lattice strain engineering plays a crucial role in adjusting the electron configuration of metal sites and regulating the interaction between the catalytic surface and adsorbate molecules, which can effectively enhance the inherent activity of the MnO2 catalyst. In this study, we report CrSb–MnO2 nanosheets with lattice strain using a phonochemical method. Doping with high-valence atoms not only enhances the electrical conductivity of MnO2 but also effectively balances the binding energy between the catalytic site and the oxygen intermediate. Additionally, it prevents the charge disproportionation reaction of Mn3+ at high currents, thereby improving the stability of the catalyst. The CrSb–MnO2 catalyst shows excellent OER activity and stability at high temperatures and high current densities under alkaline conditions. This study provides valuable insights into the design of alkaline OER catalysts. |
Manganese dioxide (MnO2) has gained significant interest as an oxidation catalyst in alkaline water due to its high density of defect edges and its ability to exist in various valence states. Additionally, according to the OER volcano diagram for metal oxides, the MnO2 catalyst exhibits activity comparable to that of noble metals.8 α-MnO2 features a large tunnelling structure that enhances the diffusion of reactant molecules, resulting in superior electrocatalytic activity compared to other crystalline forms of MnO2.9 Ma et al. reported that NiFe@α-MnO2 nanosheets achieved an overpotential of 310 mV at a current density of 100 mA cm−2 in a 1 M KOH solution. The catalytic stability was maintained for approximately 100 hours at a current density of 0.2 A cm−2.10 Additionally, Kim et al. found that MnO2 (Pruatronic®) electrocatalysts exhibited strong performance in alkaline water electrolysis (AWE), reaching 2.4 V at 1 A cm−2 with a loading of 2.5 mg cm−2 in a 1 M K2CO3 solution.11 Although the performance is relatively outstanding, it has not reached the practical application requirements of AWE electrolytic cells (1.8 V@1 A cm−2).12 At high current densities (greater than 0.5 A cm−2), Mn ions in MnO2 are prone to overoxidation during operation. This process can lead to the formation of manganese tetraoxide (MnO42−) or permanganate (MnO4−). Such overoxidation results in the loss of catalyst mass and a reduction in catalytic activity.13 Furthermore, the inherent OER activity and low electrical conductivity of transition metal oxides limit their effectiveness as catalysts in water oxidation.
Lattice strain engineering improves catalyst activity and stability by introducing vacancy defects or by doping with various metal ions or non-metal components.14 These modifications can alter metal coordination bonds through either stretching or compression. Recently, metal ions such as Mg, Fe, Co, Ni, Cu, Mo, and Ru have been identified as effective dopants for α-MnO2 catalysts.15,16 Recently, researchers have successfully developed IrO2/α-MnO2 and γ-MnO2 catalysts that employ lattice strain-engineering. These catalysts show that interfacial lattice mismatch can enhance both the OER activity and stability in proton exchange membrane (PEM) systems.17,18
Doping transition metal oxides with high-valence metal ions improves their conductivity, which helps reduce energy loss in electrolyzers. The high-valence states of these metal ions enhance the electronic configuration of the eg orbitals, effectively balancing the binding energy between catalytic sites and oxygen intermediates. As a result, this enhancement leads to improved catalytic performance.19 For example, incorporating Cr atoms into MnFeCo alloys has been shown to enhance overall catalytic performance, resulting in a low overpotential of 295 mV at a current density of 100 mA cm−2. According to d-band theory, the tensile strain introduced by adjacent Cr raises the d-band center of the active metal, which in turn affects the chemisorption of intermediates during the OER.20 Nonetheless, the application of Cr-doped MnO2 catalysts, utilizing lattice strain effects, to enhance catalyst performance in AWE remains limited. Additionally, achieving high stability in electrolyzers at elevated current densities presents a challenge due to the unavoidable phase separation between high-valent dopants and 3d metal sites. Bimetallic-doped catalysts demonstrate enhanced catalytic performance compared to those with a single metal ion. This improvement is largely attributed to the limited ability of single-doped catalysts to optimize OER activity. By altering the distribution of local electron density, bimetallic ions can enhance the synergistic and electronic effects among the different metals involved, thereby improving the stability and reaction rates of intermediates in catalytic processes.21
In this study, we synthesized α-MnO2-based catalysts using a phonochemistry method followed by high-temperature calcination. The phonochemistry technique utilizes high-frequency ultrasound to induce cavitation bubbles, resulting in the formation of nanoscale particles with associated lattice defects (Fig. 1a).22 We leveraged lattice strain to improve the performance of the α-MnO2 catalyst in alkaline water oxidation. The CrSb–MnO2 catalyst, with a Cr loading of about 2 wt% and an antimony (Sb) loading of about 1 wt%, demonstrated an overpotential of 263 mV at a current density of 100 mA cm−2 (240 mV@10 mA cm−2) in a 1 M KOH solution, which is better than most of the reported MnO2-based catalysts (>300 mV) (Fig. S1a, ESI†). Notably, this catalyst achieved an impressive performance of 1.69 V at 1 A cm−2 in alkaline water electrolysis and maintained stability at a large current of 1 A cm−2 for at least 100 h, outperforming most non-precious metal catalysts (Fig. S1b, ESI†). In situ measurements of DEMS and FTIR confirmed the role of the key reaction intermediate (*OOH) in the alkaline water oxidation process involving the CrSb–MnO2 catalyst. Furthermore, it was demonstrated that the catalytic process follows the adsorbate evolution mechanism (AEM), which helps maintain stability under high current conditions in alkaline media.
According to Bragg's rule, the lattice parameters of the host expand when subjected to high-temperature calcination or when doped with high-valence metal atoms or with metal dopants that have a larger atomic radius.23,24 Fig. 2a shows that the XRD pattern of α-MnO2 matches the standard phase, with no additional peaks detected, indicating the absence of impurity phase formation after doping. We observe that the (211) planes shift towards a lower angle. This indicates an expansion of the lattice parameters of α-MnO2 in the c direction.23 When 2.2 wt% of Cr and 1.1 wt% of Sb are added to α-MnO2 (ICP determined the content, Table S1, ESI†), the diffraction peak observed at 36.26° shifts to a lower angle by 0.15°. Similarly, with the inclusion of 2.5 wt% Cr and 1.3 wt% Co, this diffraction peak shifts slightly by 0.09° towards a lower angle. In contrast, when Fe, Co, and Ni are used as dopants, no significant shift in the diffraction peaks is observed. This indicates that introducing transition metals with larger atomic radii will efficiently generate tensile strain of the host MnO2, because the doped M–O bond length can change the ortho Mn–O bond length.25,26
![]() | ||
Fig. 2 (a) XRD patterns of bimetallic ion-doped α-MnO2 and bare α-MnO2. (b) and (c) TEM images of CrSb–MnO2 and CrCo–MnO2. (d) IFFT images of α-MnO2, CrSb–MnO2, and CrCo–MnO2. |
The atomic-resolution transmission electron microscopy (TEM) techniques enable precise measurement of the lattice spacing of solid surfaces. As illustrated in Fig. 2b and c and Fig. S2 (ESI†), we obtained a clear distribution of the lattice fringes for the CrSb–MnO2 and CrCo–MnO2 catalysts, which allowed us to quantify the strain resulting from doping. We performed selected area electron diffraction (SAED) on specific regions of the TEM images, as illustrated in the insets of Fig. 2b and c. This analysis revealed that both CrSb–MnO2 and CrCo–MnO2 nanosheets exhibit a polycrystalline structure. Consistent with previous studies,27 we focused on the most exposed (211) crystalline surface to analyze the lattice strain effects of the α-MnO2 catalysts. The measured lattice spacings (d) for the (211) crystalline surfaces of the α-MnO2, CrCo–MnO2, and CrSb–MnO2 catalysts were 1.99 Å, 2.01 Å, and 2.08 Å, respectively (Fig. 2d). The increase in the d value indicates that tensile strain forms after doping. This boosts the transition-metal d-band center and raises the antibonding d state, thereby strengthening the interaction between the active sites of the catalyst and reaction intermediate, which ultimately enhances electrocatalytic activity.14 According to the equation in the literature (see ESI†),28 we can evaluate the tensile strain intensity present within or on the surface of the catalyst. The incorporation of bimetallic ions modifies the metal coordination bonds, leading to tensile strains of 4.5% for the CrSb–MnO2 catalyst and 1% for the CrCo–MnO2 catalyst, respectively. This observation further supports that optimizing the surface geometry and electronic structure by adjusting the lattice spacing is possible. The XRD and TEM characterization clearly demonstrate the lattice strain effect in these samples, as the lattice distortion increases and the d-band center shifts away from the Fermi energy level.29 Energy dispersive spectroscopy (EDS) analysis showed a diverse elemental composition and a uniform distribution of the CrSb–MnO2 and CrCo–MnO2 catalysts (Fig. S3 and Tables S2 and S3, ESI†). These findings are consistent with the results obtained from the ICP analysis.
The electronic structure and chemical valence states of elements in bimetallic-doped catalysts were analyzed using X-ray photoelectron spectroscopy (XPS) (Fig. 3a). In the Mn 2p spectrum, the area ratio of Mn3+ to Mn4+ in α-MnO2 is approximately 0.35, indicating that defects contribute to the presence of Mn3+ in α-MnO2.30 The ratio of Mn3+ to Mn4+ in CrCo–MnO2 is approximately 0.85, whereas in CrSb–MnO2, this ratio is about 3.2. This indicates that the concentration of Mn3+ ions in CrSb–MnO2 has increased, possibly due to a higher number of oxygen vacancies.31 This result enhances the catalyst's adsorption capacity for intermediates in alkaline water oxidation.32,33 Additionally, it has been observed that the Mn 2p spectrum of CrSb–MnO2 and CrCo–MnO2 is shifted to lower binding energies compared to that of α-MnO2. This finding aligns with XPS results reported in the literature and can be attributed to the tensile strain introduced by the doping process with larger metal radius atoms.34 The Cr 2p spectra of the Cr-doped catalysts reveal the presence of Cr6+ at 579.5 eV and Cr3+–OH at 577.3 eV.35 The peak area percentages for CrSb–MnO2, are approximately 88.8% for Cr6+ and 11.2% for Cr3+–OH, while the CrCo–MnO2 catalyst shows relative peak area percentages of 55.6% for Cr6+ and 44.4% for Cr3+–OH. This indicates that the CrSb–MnO2 catalyst contains a higher proportion of high-valent Cr ions (Fig. 3b). Based on the previous reports, non-3d high-valence transition-metal ions, particularly the 6+ metal ions, can effectively tailor electronic configurations and optimize adsorption energy values of 3d transition-metal-based electrocatalysts.36,37 Additionally, Fig. 3d and Fig. S4 (ESI†) illustrate that the Sb ions in CrSb–MnO2 exist as Sb(III) and Sb(IV), and Co ions in CrCo–MnO2 are found in the trivalent oxidation state. Based on these results, the introduction of bimetallic atoms modifies the electronic structure of MnO2, resulting in a distortion of its lattice structure. This distortion facilitates an increased flow of electrons from the bimetallic atoms to the Mn ions, which in turn raises the concentration of Mn3+ ions. A higher concentration of Mn3+ ions is essential for improving the performance of the OER.38,39
As illustrated in Fig. 3d, the high-resolution XPS spectrum of the O 1s region consists of three distinct components: 529.8 eV for lattice oxygen (M–O), 531.5 eV for oxygen vacancies (Ov), and 533.4 eV for surface-adsorbed oxygen (Oa).40 The peak area percentages for M–O bonds, Ov, and Oa in CrCo–MnO2 are 82.3%, 4.4%, and 7.3%, respectively. In contrast, the proportions of M–O and Ov in CrSb–MnO2 increased to 84.6% and 13.0%, respectively, and that of the Oa decreased to 2.4%. The content of Ov in the CrSb–MnO2 and CrCo–MnO2 catalysts was significantly higher than that in the α-MnO2 catalysts. This aligns with studies in the literature showing that high-valence metal ions can increase the concentration of oxygen vacancies.41 The oxygen vacancies in metal oxides play a crucial role in optimizing the electronic structure of the metal surface and enhancing the intrinsic activity of the reactive sites.42 Electron paramagnetic resonance (EPR) studies further confirmed the presence of oxygen vacancies (Ov) in the catalysts (Fig. 3e). The results indicate that MnO2 synthesized using the phonochemical method produced a higher number of Ov compared to MnO2 prepared using the stirring method. This difference can be attributed to the addition of Tween 85 with the ultrasound effect.43
The electrochemical structure and composition of the CrSb–MnO2 and CrCo–MnO2 catalysts were analyzed using XPS after completing a stability test (Fig. 4f). In the case of CrCo–MnO2, the proportion of Mn3+ decreased from 56% to 31% before and after water oxidation. Meanwhile, the proportions of Mn4+ and Mn2+ increased by 7% and 32%, respectively. In contrast, CrSb–MnO2 exhibited a decline in Mn3+ from 83% to 75%, while the proportion of Mn4+ increased by 8%, and no Mn2+ was detected. Besides, Cr3+ in the CrSb–MnO2 catalyst after the stability experiment was readily oxidized to Cr6+ during the OER process (Fig. S11, ESI†).54 The oxidation process played a crucial role in regulating the valence state of the active metal sites, which in turn facilitated the adsorption and desorption of intermediates involved in the OER. In MnO2-based catalysts, the breaking of the Mn–O bond aids in dissolving Mn3+, leading to the formation of either Mn2+ or Mn4+, which contributes to structural collapse during water oxidation.55 The results indicated that the doping of Cr inhibited the peroxidation of Mn3+, a key factor in the water oxidation process.
To further evaluate the performance of the MnO2-based catalyst in an industrial operating system, the CrSb–MnO2 catalysts were tested in the AME water electrolyzer device. Each point on the polarization curves was obtained from measurements taken at constant voltage. As illustrated in Fig. 4c, the CrSb–MnO2 cell achieved a current density of 1.0 A cm−2 at an applied voltage of 1.69 V (1.99 V@4.43 A cm−2), surpassing the CrCo–MnO2 cell, which had a voltage of 1.73 V under the same current density (1.99 V@3.69 A cm−2). The enhancement was especially larger in the high current density region. The CrSb–MnO2 anode shows improved performance compared to recently reported transition metal-based electrodes (Fe, Co, and Ni) such as FexNiyOOH, NiFeAl and CoCrOx and is comparable to noble-metal-based anion exchange membranes such as IrO2/Ni foam (Table S4, ESI†). The ohmic resistance in the AWE electrolyzer was reduced in the CrSb–MnO2 catalyst compared to CrCo–MnO2 catalysts, indicating that the enhanced electrical conductivity of the CrSb–MnO2 catalyst improved the PEM water electrolyzer performance (Fig. S12, ESI†).56 The long-term stability performance of CrSb–MnO2 was assessed, as the stability of the catalyst is vital for its practical applications. The stability tests for CrSb–MnO2 and CrCo–MnO2 catalysts, using the constant voltage method, were conducted under harsh industrial conditions. The CrSb–MnO2 cell exhibited remarkable stability during continuous operation for 100 h at a current density of 1 A cm−2, showing no significant degradation or reduction in activity (Fig. 4d and Table S4, ESI†). In contrast, the CrCo–MnO2 cell experienced a notable decrease in performance after just 60 h of continuous operation (Fig. S13, ESI†). The ICP-MS results of the electrolyte after a stability test in 1 M KOH indicated that the dissolution of Mn ions in the CrSb–MnO2 catalyst was lower than that observed in previously reported Mn-based catalysts (Tables S5 and S6, ESI†). These results demonstrate a substantial increase in the activity and stability of the CrSb–MnO2 electrodes under harsh industrial conditions, making them suitable for AWE applications.
MnO2-based catalysts facilitate water oxidation primarily through two mechanisms: the adsorbate evolution mechanism (AEM) and the lattice oxygen oxidation mechanism (LOM) (Fig. S14, ESI†). Research has shown that the AEM pathway, which involves the *OOH intermediate, is more effective than the LOM pathway that includes the *O–*O intermediate, in enhancing the stability of MnO2 catalysts. The *O–*O intermediates can create vacancies in the lattice structure, which may ultimately lead to the collapse and dissolution of the catalyst, posing challenges for its stability.63 The AEM mechanism is preferred for the oxygen evolution process on the (211) crystal surface, as it helps maintain the stability of the crystal structure, even under elevated anode potentials.64
In situ FTIR spectroscopy was used to investigate the mechanism of the OER and to identify the intermediates formed during the electrocatalytic process. These tests were conducted at various potentials, ranging from the open-circuit potential (OCP) up to 1.6 V (vs. RHE). In Fig. 6a, for the CrSb–MnO2 system, as the applied potential increases, the driving force on the electrode surface also increases. This enhances the generation rate of the intermediates *OH and *OOH, leading to a higher concentration of these intermediates. Consequently, a distinct absorption peak is observed at approximately 1240.9 cm−1. The O–H vibration peak appeared at 3314.6 cm−1, which was attributed to the characteristic vibrational adsorption of the *OOH intermediate.65 As shown in Fig. 6b for α-MnO2, a peak for the *OOH intermediate is observed at approximately 1240.9 cm−1. Additionally, a signal peak corresponds to the *O–*O intermediate, which appears around 1056 cm−1. The production of *O–*O intermediates is associated with the lattice oxygen oxidation mechanism (LOM), and this peak becomes more pronounced as the potential increases.61 These indicate that the AEM and the LOM are involved in the OER process of α-MnO2. It has been demonstrated that the release of lattice oxygen through the LOM mechanism accelerates the dissolution and overoxidation of α-MnO2.18 The instability of α-MnO2 catalysts was linked to lattice oxygen involvement. Thus, CrSb–MnO2 was primarily characterized by AEM during the OER process, demonstrating its effectiveness in suppressing the LOM pathway.
Using 18O isotope-labeled differential electrochemical mass spectrometry (DEMS), we further verify the OER process of CrSb–MnO2 preferred AEM, which restrains the overoxidation and dissolution of Mn species. As shown in Fig. 6a and b, mass signals corresponding to 34O2 (16O–18O) and 36O2 (18O–18O) were detected for the CrSb–MnO2 catalysts during the OER process. The signal for 36O2 originates from the 18O2 in the electrolyte, while the 34O2 signal arises from the lattice oxygen in the oxide catalyst that has not been isotopically substituted.59,66 For the α-MnO2 catalyst the mass signals for 32O2 and 34O2 were detected. The signal of 32O2 (16O–16O) indicates that 16O in the lattice is directly involved in the production of O2, and if the lattice oxygen combines with 18O in the electrolyte then the signal of 34O2 is produced.59,67
We also performed a quantitative analysis to compare the percentage of oxygen originating from the lattice versus that from the electrolyte. The results showed that the proportion of lattice oxygen in CrSb–MnO2 (16.2%) was significantly lower than that in α-MnO2 (64.8%). This indicates that CrSb–MnO2 favors oxygen evolution through the AEM rather than the LOM, as illustrated in Table S7 (ESI†).68,69 After cleaning the catalyst surface using H216O, we placed electrodes with the α-MnO2 and CrSb–MnO2 catalysts into the H216O electrolyte. During the oxygen evolution process, we detected a strong signal for 32O2 and a very weak signal for 34O2 (16O–18O) from both the α-MnO2 and CrSb–MnO2 catalysts. Notably, there was no signal for 36O2 (18O–18O) observed (Fig. 6e and f). Following quantitative calculations, we found that the ratio of 34O2 (16O–18O) to 32O2 (16O–16O) in CrSb–MnO2 (21%) is significantly lower than that in α-MnO2 (37%). This preference contributes to the enhanced stability of the oxygen evolution at high current densities.
The theoretical calculations suggest that CrSb–MnO2 is the optimal configuration for AEM when compared to Cr–MnO2 and α-MnO2. This configuration effectively lowers the energy barrier for the adsorption of *O and facilitates the rapid formation of *OOH in AEM. The improvement is attributed to the combined effects of electronic interactions and tensile strain present in the bimetallic doped Mn-based catalysts. These factors reduce the binding energy of *O and *OOH at the active sites, resulting in faster reaction kinetics.70–72 At a voltage of U = 0 V, all reaction steps are endothermic processes (Fig. S15, ESI†). In contrast, the step from *O to *OOH at U = 1.23 V (Fig. 7b) is identified as the potential-determining step (PDS) for the OER. This step is characterized by having the maximum energy difference (ΔG) among all intermediate conversions that occur during the alkaline OER.70,73 Notably, the *O formation step on CrSbO2 exhibits the lowest energy barrier at −0.12 eV when compared to Cr–MnO2 and α-MnO2. This suggests that the process is advantageous for the rapid formation of *OOH.70 As illustrated in Fig. 7c, CrSb–MnO2 shows the lowest free energy barrier compared to Cr–MnO2 and α-MnO2, indicating that it requires the least potential to drive the OER. These findings highlight that incorporating Cr and Sb into the MnO2 lattice significantly enhances OER activity.74
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ey00106d |
This journal is © The Royal Society of Chemistry 2025 |