Tianzhen
Ren‡
a,
Lu-Kang
Zhao‡
a,
Xiao
Zhang‡
a,
Xuan-Wen
Gao
*ab,
Hong
Chen
a,
Zhaomeng
Liu
*a,
Dongrun
Yang
a and
Wen-Bin
Luo
*a
aInstitute for Energy Electrochemistry and Urban Mines Metallurgy, School of Metallurgy, Northeastern University, Liaoning 110819, China. E-mail: luowenbin@smm.neu.edu.cn
bKey Laboratory of Adv. Energy Mater. Chemistry (Ministry of Education), Nankai University, Tianjin 300071, China
First published on 10th April 2025
Perovskite oxides, with their flexible electronic structure and low cost, are highly attractive alternatives to noble metal catalysts for oxygen redox reactions (ORRs). Herein, we report LaNiO3 perovskite as an efficient ORR catalyst, where oxygen vacancies and oxygen-containing functional groups work synergistically. LaNiO3 was subjected to thermal shock in different media to introduce defects on the structural surface. Experimental results indicate that thermal shock in air creates oxygen vacancies on the catalyst surface, while thermal shock in aqueous solution increases the content of oxygen-containing functional groups (–OH) on the surface. The samples subjected to thermal shock exhibit superior ORR catalytic performance, with the limiting current density increasing from 4.2 mA cm−2 to 5.4 mA cm−2. This work provides a convenient and straightforward approach for constructing a series of materials enriched with surface defects.
ORR is a complex process involving multiple steps in aqueous solution, such as dissociation, adsorption, and electrochemical reactions, leading to the formation of oxygen-containing species like OH−, O2−, and HO2−.20,21 Extensive studies have demonstrated that the effective adsorption of oxygen on the electrode surface, the dissociation of the O–O bond, and the extraction of oxide species significantly influence the overall kinetics of the ORR.22 The ORR can be regarded as an ensemble of oxygen exchange processes occurring on the catalyst surface. According to the Sabatier principle, catalytic activity can be modulated by altering the bond strength between the catalyst surface and the reactants/products.23 However, this is merely a qualitative guideline, as the adsorption strength of oxygen on the catalyst is largely dependent on the various methods employed for catalyst preparation. From the perspective of molecular orbital theory, the hybridization of the d orbitals of the B-site metal and O 2p orbitals forms the antibonding σ* (eg) and π* (t2g) orbitals. The literature indicates that the electron filling of the eg orbital of the B-site cations is directly related to the catalytic activity of ABO3 perovskite oxides.1,12,24 It is worth noting that, compared with cation doping, the introduction of oxygen vacancies or surface engineering has been considered a simpler and more effective method to enhance the activity of perovskite oxides recently.21,25,26 This approach not only adjusts the hybridization between the transition metal 3d and O 2p orbitals but also, with appropriate surface engineering, enhances the conductivity and activates the adsorbed O2.
Previous studies have identified LaNiO3 as one of the most promising catalysts for the ORR;27–29 however, its reaction kinetics remain severely limited. Herein, we report a strategy to significantly enhance the ORR electrocatalytic activity of LaNiO3 (LNO) through thermal shock in different media, without any additional modifications. Among LNO samples subjected to thermal shock in various media, LNO-W (quenched in deionized water) exhibited the best ORR electrocatalytic performance. Compared with LNO-F (cooled in furnace naturally), quenching in air and solution increases the content of surface functional groups and oxygen vacancies. We further explored the mechanisms underlying the enhanced ORR catalytic activity.
![]() | ||
Fig. 1 (a–d) Rietveld refinement of XRD patterns, (e) lattice parameter for LNO-F, LNO-A, LNO-F, and LNO-N. The inset shows the crystal model of LaNiO3. |
The morphology of LNO samples was analysed by SEM and TEM. As shown in Fig. 2, all of LNO-F, LNO-A, LNO-W, and LNO-N exhibit irregular block-like structures, which are composed of nanosized particles with diameters ranging from approximately 40 to 160 nm. Notably, the particles of the samples subjected to thermal shock exhibited varying degrees of surface cracking, which promote O2 transport in samples. Furthermore, Fig. 2b, d, f and h show locally magnified SEM images for LNO-F, LNO-A, LNO-W, and LNO-N, which show that the particles are more dispersed after thermal shock and more surfaces are exposed, especially in solution. The histograms presented in the inset show the particle size distributions obtained by counting at least 100 particles in SEM images. The result shows a significant decrease of particle size after thermal shock. Notably, the catalyst particles exhibited greater interparticle distances when subjected to thermal shock in solution, which may be associated with the etching effect of OH− ions. This is expected to increase the specific surface area, and facilitate exposure of more active sites.5,29,31
![]() | ||
Fig. 2 (a, c, e and g) SEM micromorphology images, and (b, d, f and h) partial enlarged details, inset shows grain size distribution of LNO-F, LNO-A, LNO-W, and LNO-N. |
Further insights into the microstructure of nanosized particles were investigated through TEM. Apparently, when subjected to thermal shock in different media, the particle morphology undergoes continuous changes. LNO-F and LNO-A particles exhibit regular hexagonal shapes, while the particles of LNO-W and LNO-N gradually become smooth (Fig. S2†). Notably, thermal shock not only induced cracks on macroscopic particles but also created fissures within nanoparticles, which is attributed to the refinement of particle size.
As shown in Fig. 3, the LNO-F particles exhibit good crystallinity both on the surface and in the bulk. In contrast, the particles of the samples subjected to thermal shock consist of well-developed crystals and a disordered surface layer. The result probably originated from the successful introduction of oxygen vacancies leading to the disintegration and discontinuous surface lattice structure.32 Moreover, the grains of LNO-A, LNO-W, and LNO-N contain a large number of structural defects, such as dislocations and stacking faults, as shown in Fig. 3c, f and i. The Geometric Phase Analysis (GPA) of the lattice reveals evident stress concentration at the locations of these structural defects. Stress concentration is a significant factor leading to grain refinement and changes in crystal orientation. Therefore, the particle diameter of the samples after thermal shock is significantly reduced, and the particles of LNO-W and LNO-N evolve into those composed of elongated subgrains (Fig. S2† and 3g). As shown in Fig. 3b and d, the lattice spacing corresponding to the (101) plane in LNO-A (0.398 nm) is significantly larger than that in LNO-W (0.386 nm), which is consistent with the Rietveld refinement of XRD results. The formation of oxygen vacancies increases the repulsion between cations, leading to cell expansion. However, on quenching in deionized water, some of the oxygen vacancies are occupied by oxygen-containing functional groups, which in turn mitigates the lattice expansion.21
The ORR involves a series of processes in which oxygen species are exchanged on the catalyst surface. To further investigate the characteristics of surface elemental composition and electronic structure of LaNiO3 perovskite catalysts subjected to thermal shock in different media, XPS analysis is carried out. The survey spectra, shown in Fig. S3a,† indicate that all samples have the same elemental composition with no detectable impurities on the surface. To determine whether N element is present in the LNO-N catalyst, we performed XPS analysis on the N 1 s spectrum of LNO-N (Fig. S3b†). The results revealed negligible presence of N element on the sample surface. Therefore, we focused solely on the influence of the hydroxyl concentration in the ammonia solution on the experiments. The O 1s XPS spectra consist of four distinct peaks at 528.3 eV, 529.8 eV, 531.1 eV, and 533.1 eV, corresponding to LO, oxygen vacancy (VO), adsorbed oxygen, and H2O, respectively (Fig. 4a and S4a†).9,11,33,34 A significant increase in the content of oxygen vacancies was observed in LNO-A, indicating that a certain amount of oxygen defects had been effectively incorporated into the surface of LNO-A nanoparticles. Moreover, reducing the quenching temperature (LNO-A-700) results in a marked decrease in the proportion of oxygen vacancies on the sample surface. This suggests that oxygen vacancies are generated on the sample surface at high temperatures, and quenching in air does not provide sufficient time for these vacancies to recombine, leading to their retention.35 Some studies have shown that the adsorption of HO2− by the oxygen vacancies and transition metal can reduce the dissociation of HO2−, which is conducive to the further reaction to form OH− and complete the 4 e− reaction.21 Notably, the proportion of adsorbed oxygen species (e.g., hydroxylation) on the sample surface increased in LNO-W, while the proportion of oxygen vacancies decreased (Fig. S4b†). This may be due to H2O molecules or OH− in the solution occupying the positions of oxygen vacancies during the quenching process. It is noteworthy that for samples subjected to thermal shock in solution, the peak corresponding to adsorbed oxygen shifts towards higher binding energies, indicating an enhanced adsorption capacity for oxygen on the sample surface.36 According to studies on the mechanism of the ORR, adsorbed oxygen is considered to be the most active site for the ORR.10,34,37 Therefore, increasing the concentration of hydroxyl groups can effectively modulate the catalyst surface's adsorption capacity for intermediate products.
![]() | ||
Fig. 4 (a) O 1s, (b) valence band, (c) Ni 3p XPS spectra of LNO-F, LNO-A, LNO-W, and LNO-N, (d) proposed mechanisms for the ORR in an alkaline medium. |
The XPS spectra of Ni 2p and La 3d are shown in Fig. S5(a and b),† respectively. The analysis of the XPS results is quite complex because the intensity of Ni 2p is difficult to distinguish from other spectral features. The peaks of La 3p3/2 and Ni 2p3/2 overlap, as well as the Ni 2p1/2 signal loss feature belonging to La 3p3/2.18,38 Therefore, we conducted a qualitative analysis of the La and Ni elements in this part. According to the literature, the peaks at 833.7 eV and 835.5 eV in the La 3d5/2 spectrum are attributed to La2O3 (peak d) and La(OH)3 (peak c), respectively.27,38 The significant increase in the intensity of peak c in LNO-W and LNO-N indicates an enhanced degree of hydroxylation. To further analyse the variation for Ni, the XPS spectrum of Ni 3p is presented in Fig. 4c. Compared to LNO-F, the Ni 3p peak of samples subjected to thermal shock shifts towards higher binding energies. Particularly in LNO-W and LNO-N, the Ni 3p peak undergoes a significant leftward shift, consistent with the trend observed for the Ni 2p peak. This indicates an increased proportion of Ni3+ in the samples. The presence of more empty orbitals in Ni3+ effectively facilitates electron transfer, thereby accelerating the ORR process. Moreover, the higher electronegativity of Ni3+ enables efficient adsorption of intermediates, enhancing the ORR performance.39
The electronic structure near the top of the valence band is significant for ORR performance. The valence band spectra of the perovskite samples cooled in different media are shown in Fig. 4b. Generally, valence band spectra can be divided into four parts, and associated with the contributions of hybridized states, including Ni 3d–O 2p, O 2p non-bonding states, Ni t2g states, and Ni 3d eg states.18,23 The ORR process involves the transformation of various intermediates, with the rate-limiting step being the displacement of surface hydroxyl groups (highlighted in red in Fig. 4d). The kinetics of this step are determined by the driving force of electron transfer from the eg orbitals to the O–O antibonding orbitals. It is observed that an increase in the proportion of oxygen vacancies leads to a shift of the O 2p non-bonding states towards the Fermi level in LNO-A, thereby reducing the charge transfer gap in the catalytic process. Moreover, the rise in the proportion of oxygen vacancies results in a decrease in the electron density of O 2p and Ni t2g, while enhancing the hybridization degree of Ni 3d–O 2p orbitals. This, in turn, causes a reduction in the Ni–O–Ni bond angle, which is consistent with the results of Rietveld refinement of XRD.31,40 Notably, the changes in valence electrons in LNO-N are exactly opposite to those in LNO-A, indicating that oxygen vacancies and hydroxyl functional groups play different roles in ORR catalysis.
Electrochemical ORR properties of LNO-F, LNO-A-700, LNO-A, LNO-W, and LNO-N were characterized. Fig. S6† shows the ORR activity of the pristine nanoparticles and nanoparticles subjected to thermal shock at different rotation speeds. Based on previous experimental conclusions, acetylene black (AB) only serves to improve the electrical conductivity of the active material and has minimal impact on the ORR activity. We used linear sweep voltammetry (LSV) based on the rotating ring-disk electrode (RRDE) to represent ORR activity and evaluated ORR performance using limiting current density (jm), onset potential (Eonset), and Tafel slope. The Eonset was defined as the potential at which the ORR current density reaches 0.1 mA cm−2. At a rotation speed of 1600 rpm and a scan rate of 10 mV s−1, the ORR LSV curves (Fig. 5a) were measured between −0.7 and 0 V vs. Ag|AgCl (3 M KCl). It clearly demonstrates that the ORR activity of the LaNiO3 powders subjected to thermal shock was significantly enhanced, as evidenced by their more positive Eonset and increased jm (Fig. 5d). Among the particles subjected to thermal shock, LNO-W showed the best ORR activity, with the half-wave potential increasing from 0.62 V to 0.72 V vs. RHE. In particular, the onset potentials of pristine LNO-F, LNO-A-700, LNO-A, LNO-W, and LNO-N were 0.787 V, 0.793 V, 0.799 V, 0.823 V, and 0.815 V vs. RHE, respectively, with LNO-W exhibiting the highest jm (5.42 mA cm−2). To gain further insights into the observed electrochemical responses, we analysed the ORR process using Tafel slopes, which provide a kinetic perspective. For LNO-F, the ORR Tafel slope decreased from 93.1 mV dec−1 to 82.2 mV dec−1, indicating superior kinetics for LNO-W compared to LNO-A and LNO-F. The Koutecky–Levich equation is used to evaluate the number of electrons transferred during the ORR process.10,32 The electron transfer number of pristine LNO-F is approximately 3.63, while the electron transfer number of LNO-W increased to 3.98, which is very close to that of commercial Pt/C.32 This suggests that LaNiO3 subjected to thermal shock tends to experience a four-electron transfer reaction, reducing oxygen to an oxygen-containing material. The significant increase in the electron transfer number after thermal shock indicates that oxygen vacancies in conjunction with hydroxylation can facilitate a four-electron reaction.
![]() | ||
Fig. 5 ORR performance of LNO-F, LNO-A, LNO-W, and LNO-N. (a) LSV curves, (b) Tafel slope, (c) Koutechy–Levich (K–L) plots, (d) onset potential and electron transfer number. |
The results clearly demonstrate that thermal shock treatment significantly enhances the ORR catalytic activity of LaNiO3 nanosized powders, representing an effective approach to enhancing the electrocatalytic oxygen reduction efficiency of LaNiO3. We attribute this enhancement primarily to the introduction of oxygen vacancies and oxygen-containing functional groups on the particle surfaces, which subsequently enhance intrinsic properties such as oxygen exchange kinetics.17,41 Notably, the overall ORR electrocatalytic performance of LNO-W surpasses that of LNO-A and LNO-N, indicating that the enhanced electrochemical performance of LNO-W results from a delicate balance between oxygen vacancies and oxygen-containing functional groups (hydroxylation). The increase in oxygen vacancy concentration leads to a shift of the O 2p non-bonding states towards the Fermi level, effectively reducing the charge transfer energy gap and enhancing the hybridization degree of Ni 3d–O 2p orbitals, thereby improving ORR performance. Conversely, an increase in the number of hydroxyl functional groups on the perovskite surface can effectively raise the concentration of adsorbed oxygen, which typically acts as the active species in the ORR process. However, hydroxyl functional groups tend to decrease the hybridization degree of Ni 3d–O 2p orbitals and enlarge the charge transfer energy gap. Therefore, there exists a synergistic effect between oxygen vacancies and oxygen-containing functional groups, which collectively optimize the ORR process. And there is an optimal surface oxygen vacancy and oxygen-containing functional group concentration that enhance oxygen exchange kinetics. These defects have a significant impact on the electronic structure of LaNiO3 surface, leading to an increased number of active sites. Moreover, the incorporation of defects induces changes in the oxidation states of cations, which increases the electrical conductivity of the LaNiO3 catalyst.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta01346a |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |