Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Supercritical CO2-modulated phase transition in CaSnO3 from orthorhombic to cubic symmetry for room-temperature ferromagnetism

Lei Xu a, Qun Xu *a and Buxing Han b
aHenan Institute of Advanced Technology, Zhengzhou University, Zhengzhou 450003, P. R. China. E-mail: qunxu@zzu.edu.cn
bInstitute of Chemistry, Chinese Academy of Science, Beijing 100080, P. R. China

Received 2nd July 2025 , Accepted 28th September 2025

First published on 15th October 2025


Abstract

The advancement of nanotechnology has enabled magnetic nanomaterials to exhibit remarkable potential and application value in medicine, transportation, information storage, and spintronics owing to their unique physicochemical properties. In this study, supercritical carbon dioxide (SC CO2) was used to successfully induce room-temperature ferromagnetism in CaSnO3 without magnetic element doping, achieving a maximum saturation magnetization of 0.0727 emu g−1 at 16 MPa. The SC CO2 treatment introduced lattice-scale defects, releasing residual force within distorted SnO6 octahedra, which led to the suppression of structural distortion and drove a structural phase transition from orthorhombic to cubic. Additionally, the enhanced symmetry was accompanied by anisotropic lattice expansion and tensile strain, which thermodynamically lowered the oxygen vacancy formation energy, thereby kinetically driving the creation of more defects. This disrupted the intrinsic antiferromagnetic order and significantly enhanced ferromagnetism. This work elucidates a defect-strain synergy mechanism for tuning material magnetic order, distinguishing it from conventional stoichiometric doping strategies and highlighting the critical role of SC CO2 in material modification.


Introduction

With the advancement of nanotechnology, magnetic nanomaterials have shown great potential and application value in areas such as medicine,1 transportation,2 information storage3 and spintronics4–7 due to their unique physicochemical properties. Room-temperature ferromagnetic materials offer great potential for a variety of advanced technology applications, as well as new possibilities for the next generation of low-power, high-density electronic and information storage devices.8 Perovskites are excellent candidates for ferromagnetic materials, benefiting from their tunable crystal structures9 and good electron correlation properties,10 as well as a strong spin–orbit coupling effect.11 Among the plethora of perovskite materials, alkaline earth stannate CaSnO3 has attracted great interest from many researchers because of its wide band gap,12 high melting point,13 and excellent optical14 and semiconductor properties.12 However, comparatively limited research has been conducted on the ferromagnetism of CaSnO3. The CaSnO3 perovskite is composed of octahedral SnO6 and its vertices are linked to each other by mutual sharing of oxygen.15 In the ideal orthorhombic structure (Pbnm) of bulk CaSnO3, the Sn4+ ion occupies the center of the oxygen octahedron. Due to its d10 electron configuration, Sn4+ has no local magnetic moment, ruling out the possibility of ferromagnetism at room temperature.16

In the existing studies, the modulation of the ferromagnetism of CaSnO3 materials is mainly through elemental doping, and there is no focus on their intrinsic ferromagnetism.17 Aadil Ahmad Bhat et al.18 synthesized and tested Mn and Ce doped CaSnO3 nanomaterials using a hydrothermal method. The doping modified the internal chemical environment of the materials, and the diamagnetic CaSnO3 host perovskite transformed into a ferromagnetic material. Therefore, is there a new method that can effectively modulate the internal structure and intrinsic ferromagnetism of CaSnO3 without introducing other elements? In our previous work, SC CO2 has been used to modulate the selective breaking of covalent bonds in CaTiO3,19 thereby exposing a specific crystalline surface that exhibits excellent room-temperature ferromagnetism. In addition, it was found that SC CO2 facilitates the conversion of external pressure into chemical pressure inside BaZrO3,20 changing the chemical environment of BaZrO3 and thus, effectively modulating its ferromagnetic properties. These endeavors capitalize on the distinctive physicochemical characteristics exhibited by supercritical CO2,21–26 including exceptional diffusivity, negligible surface tension, and adjustable solvation power,27 which enable its selective penetration into crystalline matrices to facilitate defect generation and strain field redistribution while preserving the overall structural integrity of the crystalline framework.

In this study, an alternative strategy to conventional elemental doping is proposed, in which SC CO2 is utilized to induce lattice-scale defects in CaSnO3, thereby effectively releasing residual force within distorted SnO6 octahedra and driving a structural phase transition from orthorhombic to cubic symmetry. In addition, the symmetry elevation induces pronounced anisotropic lattice expansion and tensile strain, which thermodynamically lowers the oxygen vacancy formation energy, thereby kinetically driving the creation of more defects, and ultimately induces the antiferromagnetic CaSnO3 to exhibit room-temperature ferromagnetism after SC CO2 treatment.

Results and discussion

Structural characterization of CaSnO3

The structures of the CaSnO3 nanosheets prepared under different SC CO2 pressures were characterized by X-ray diffraction (XRD). Fig. 1a illustrates the XRD patterns of CaSnO3 containing labeled diffraction peak sites. Analysis of these peaks using Jade software confirmed that the untreated sample (bulk) exhibited an orthorhombic phase (PDF card no. 77-1797, space group Pbnm).28 Unexpectedly, the SC CO2-treated CaSnO3 exhibited a coexistence of Pbnm and Fm[3 with combining macron]m phases.29 During the process of conversion from Pbnm to Fm[3 with combining macron]m, the crystal structure and cell parameters of CaSnO3 change dramatically: the symmetry of the crystal is increased, and the bond angle between its two SnO6 octahedra becomes larger (from 148° to 180°). This was attributed to the anisotropic expansion of the orthorhombic lattice due to SC CO2 penetration (see Table S1, SI), which lowers the phase transition energy barrier and induces the emergence of new cubic phases.30 Specifically, the analysis focused on three characteristic peaks, (110), (112), and (004). As shown in Fig. 1b–d, the diffraction peak of CaSnO3 after SC CO2 treatment shifted towards lower angles. When the pressure of SC CO2 reaches 16 MPa, the three characteristic peaks are most clearly shifted to a lower angle, providing key evidence for the increased lattice spacing originating from the strain effect due to the SC CO2 treatment.
image file: d5sc04873g-f1.tif
Fig. 1 Characterizations of CaSnO3 treated under different SC CO2 pressures. (a) Survey characterization of XRD. Enlarged views of some planes: (b) (110) plane, (c) (112) plane, and (d) (004) plane. (e) The Rietveld structure refinement and structure under 16 MPa.

To further confirm the changes in the crystal structure of the samples after SC CO2 treatment, the XRD diffraction data were refined using Rietveld structural refinement. This analysis was conducted to assess the alterations in cell parameters and validate the formation of the new phase. The bulk CaSnO3 crystals usually tend to form orthorhombic phases due to the difficulty in releasing structural stress. When the SC CO2 pressure was increased from 14 to 16 MPa, it kinetically facilitated the increase of oxygen vacancy defects inside the crystal and prompted the release of lattice stresses around the defects, and the content of the cubic phase CaSnO3 was changed from 19% to 32% (see Fig. 1e and S1, SI). However, further pressure increase to 18 MPa unexpectedly reduced the cubic phase content to 17% (see Fig. S1, SI). This reversal could potentially originate from excessive SC CO2 infiltration, which may have driven the elastic strain energy of the cubic phase beyond its thermodynamic stability threshold. Such over penetration likely facilitated amorphous phase formation through destabilization of the crystalline lattice.

To characterize the micromorphology of the prepared CaSnO3 with different pressures of SC CO2, high-resolution transmission electron microscopy (HRTEM) is used. Fig. 2a presents the TEM image of the untreated sample, revealing a well-defined orthorhombic structure. This observation is further corroborated by the SAED pattern shown in Fig. 2e, which exhibits distinct diffraction spots corresponding to the (210) and (004) crystallographic planes. These characteristic features are consistent with the Pbnm phase structure of CaSnO3, confirming the crystallographic orientation and phase purity in the absence of SC CO2 treatment. Under a 14 MPa CO2 pressure, the TEM analysis in Fig. 2b revealed the structural coexistence of both orthorhombic and cubic CaSnO3 phases, with the corresponding SAED pattern in Fig. 2f displaying dual diffraction arrays. Crystallographic indexing unambiguously assigned these diffraction features to the Pbnm orthorhombic system and Fm[3 with combining macron]m cubic system. Similarly, Fig. 2c and g show a pair of TEM images and SAED diffraction pattern corresponding to treatment with a 16 MPa CO2 pressure, exhibiting characteristic lattice spacings of 0.362 nm and 0.286 nm corresponding to the orthorhombic (111) and cubic (220) planes. Compared to standard CaSnO3 PDF cards (Pbnm phase: PDF#77-1797; Fm[3 with combining macron]m phase: PDF#75-0463), both planes show maximum lattice stretching at 16 MPa, indicating that SC CO2 pressure-driven structural changes occur preferentially along these crystallographic orientations. However, when the CO2 pressure increases from 16 to 18 MPa, the cubic phase CaSnO3 is difficult to observe in the TEM image presented in Fig. 2d. This phenomenon likely stems from the enhanced density and solvation power of SC CO2 under elevated pressures, which intensifies interactions with the material. The cubic CaSnO3 phase, characterized by its higher crystallographic symmetry and expanded oxygen octahedral network, exhibits heightened susceptibility to structural disordering. These features lower the Gibbs free energy barriers for amorphous nucleation, facilitating phase transformation kinetics, thereby inducing localized amorphization within cubic phase domains (see Fig. S2, SI). The diminished long-range crystallographic coherence becomes challenging to resolve via TEM imaging, while short-range ordered atomic configurations remain detectable through characteristic SAED pattern in Fig. 2h, reflecting residual structural correlations at sub-nanometer scales.


image file: d5sc04873g-f2.tif
Fig. 2 Characterizations of CaSnO3 treated under different SC CO2 pressures. (a–d) High-magnification TEM images of the bulk material and that under pressures of 14, 16 and 18 MPa. (e–h) SAED diffraction patterns of CaSnO3 samples of the bulk material and those under pressures of 14, 16 and 18 MPa.

Additionally, scanning electron microscopy (SEM) and energy dispersive spectroscopy (EDS) were employed to examine the surface morphology and elemental distribution of CaSnO3 before and after treatment (see Fig. S3 and S4, SI). Significant changes in surface morphology were observed following SC CO2 treatment. EDS mapping further revealed a decrease in oxygen content with increasing pressure, providing direct evidence for the increased oxygen vacancy concentration, which is attributable to the pressure-induced penetration of SC CO2, which progressively increased lattice defects and induced surface amorphization.

Internal structures and molecular vibrations of CaSnO3 treated under different SC CO2 pressures were investigated using Raman spectroscopy, as shown in Fig. 3a. Group theoretical analysis reveals that the CaSnO3 bulk material with the Pbnm space group exhibits 24 Raman-active modes (Γ Raman = 7Ag + 5B1g + 7B2g + 5B3g).31 The samples all show three characteristic peaks in common, with spectral bands in the range 164–227 cm−1 (B2g) possibly related to octahedral rotation of oxygen,32 and bands near 270 cm−1 (Ag) and 351 cm−1 (B1g) related to shifts in Ca and O.32


image file: d5sc04873g-f3.tif
Fig. 3 Characterization of CaSnO3 treated with different pressures of SC CO2. (a) Raman spectra. Enlarged views of some characteristic peaks: (b) 176 cm−1, (c) 270 cm−1, and (d) 351 cm−1.

In particular, the bands associated with the bending or stretching of Sn–O bonds near 420 cm−1 (B2g) and the twisting of SnO6 octahedra near 580 (Ag) and 695 cm−1 (B1g)17 disappeared after treatment with SC CO2. This arises from the simplicity of the vibrational modes and activity suppression due to the elevated symmetry of the crystal.33 The orthorhombic phase has multiple structural distortions, such as tilting of the oxygen octahedra and cation displacements. These distortions induce Brillouin zone folding effects that generate multiple Raman-active modes. However, the restoration of cubic symmetry eliminates these distortions, causing originally split vibrational modes (e.g. B2g/B1g/Ag) to merge into higher-symmetry representations through mode simplification, resulting in the disappearance of some of the peaks. Specifically, the Sn–O bond bending/stretching and octahedral twisting modes become Raman-inactive under cubic symmetry constraints. Consequently, the disappearance of specific peaks in the 400–800 cm−1 range reflects a fundamental symmetry shift in lattice dynamics during the orthorhombic to cubic phase transition. In addition, the surviving Raman peaks in Fig. 3b–d exhibit systematic redshifts after SC CO2 treatment. The B2g peak shift originates from symmetry recovery in oxygen octahedra, where the elimination of rotational distortions increases Sn–O bond lengths and reduces vibrational force constants. The Ag and B1g peak displacements arise from local polarization suppression; cubic symmetry constraints force Ca2+ ions back to high-symmetry positions, annihilating displacement-dependent vibrational modes while weakening residual Ca–O coupling, leading to coordinated frequency downshift.

X-ray photoelectron spectroscopy (XPS) was employed to characterize the oxidation states of tin and oxygen. The total XPS characterization (see Fig. S5, SI) indicates that no other magnetic impurities are present in CaSnO3. Fig. 4a illustrates that the XPS of O 1s in CaSnO3 can be divided into three peaks with binding energies of 529.9, 531.7, and 533.6 eV. These peaks were assigned to the lattice oxygen OL (Sn–O–Sn and Ca–O), the oxygen vacancy OV, and oxygen adsorption OA, respectively.34 A notable transformation is observed in comparison to the bulk material, and the percentage of oxygen vacancies increases from 36% to 40%, 48%, and 39%, following SC CO2 treatment at different pressures. Fig. 4b illustrates that the XPS spectra of Sn exhibit two principal peaks, which correspond to Sn 3d5/2 and Sn 3d3/2 with binding energies of 486.4 and 494.8 eV, respectively. A shift of the two main peaks is observed in the samples treated with SC CO2, so the emergence of two additional peaks at low binding energies is evident, which can be attributed to Sn3+. Furthermore, an increase in CO2 pressure resulted in a corresponding change in Sn3+ content. The observed phenomenon originates from the chemical reaction within CaSnO3, leading to the simultaneous formation of Sn3+ species and OV. Specifically, each OV formation corresponds to the reduction of two adjacent Sn4+ cations to Sn3+ through charge compensation mechanisms.


image file: d5sc04873g-f4.tif
Fig. 4 XPS characterization under different SC CO2 pressures. (a) O 1s XPS. The percentage contents in (a) represent OV. (b) Sn 3d XPS. The percentage contents in (b) represent Sn3+. (c) Electron paramagnetic resonance (EPR) spectra.

The EPR spectra are illustrated in Fig. 4c. The samples treated with SC CO2 exhibit pronounced signals at g = 2.004 and g = 1.97, which corroborate the generation of Ov35 and Sn3+.36 The signal with g = 2.004 can be assigned to OV, and the other signal at g = 1.97 is attributed to Sn3+. As illustrated in Fig. 4c, the intensities of the characteristic peaks corresponding to OV and Sn3+ show a significant increase when the CO2 pressure is increased to 16 MPa, and then show a period of weakening as the CO2 pressure increases to 18 MPa. This phenomenon can be attributed to the fact that increased pressure in the range of 14–16 MPa enhances the penetration of SC CO2, which intensifies the disruption of Sn–O bonds and promotes the formation of OV and Sn3+. Additionally, as the CO2 pressure continues to increase, it promotes the formation of structurally disordered amorphous phases, leading to the migration or annihilation of oxygen vacancies and affecting the stability of Sn3+. This trend is consistent with the XPS results.

Magnetic properties

Fig. 5a and b show the room temperature M–H plots of CaSnO3 samples before and after SC CO2 treatment. Due to the presence of a d10 configuration, pristine CaSnO3 is assumed to exhibit diamagnetic behavior.37 The experimental results revealed that samples subjected to SC CO2 treatment demonstrated pronounced magnetic ordering, with ferromagnetic properties displaying marked variations depending on the applied CO2 pressure during processing. The saturation magnetization intensity of CaSnO3 was observed to increase to 0.0327, 0.0727, and 0.0253 emu g−1 at CO2 pressures of 14, 16, and 18 MPa, respectively. At a pressure of 16 MPa, the sample achieves the highest saturation magnetization and also reaches a maximum residual magnetization of 0.0122 emu g−1. Furthermore, this saturation magnetization intensity is higher than that observed for CaSn1−xMnxO3 with x = 0.1.18 Therefore, SC CO2 exhibited a remarkable ability to modulate ferromagnetism during the treatment of CaSnO3. As illustrated in Fig. 5d, the field cooling–zero magnetic field cooling (FC–ZFC) curve of CaSnO3 treated at 16 MPa exhibits non-overlapping behavior below 400 K. This indicates that the Curie temperature after SC CO2 treatment has exceeded 400 K, demonstrating that CaSnO3 treated with SC CO2 can maintain ferromagnetic properties at temperatures above room temperature. The 16 MPa sample demonstrates a notably enhanced Curie temperature (TC) compared to the ZFC curves of CaSnO3 treated at 14 and 18 MPa (see Fig. S6, SI). This enhancement likely originates from an increased propensity for structural phase transitions and a higher cubic phase fraction under the 16 MPa conditions. The cubic phase, characterized by its superior symmetry and optimized bond angles for effective magnetic exchange interactions, promotes the establishment of long-range magnetic ordering. Furthermore, its inherently stronger magnetic coupling strength compared to lower-symmetry phases culminates in an elevated TC at this specific pressure.
image file: d5sc04873g-f5.tif
Fig. 5 Ferromagnetic behavior. (a) M–H curves at 300 K. (b) The M–H curves near H = 0. (c) Bar graph showing specific magnetic parameters. (d) FC–ZFC magnetization curve obtained at 16 MPa in an external magnetic field of 500 Oe.

Experimental

Materials

CaSnO3 (99.9%) powder was purchased from Shanghai Aladdin Biochemical Technology Co., Ltd. It was used directly, without further purification. CO2 of 99.99% purity was purchased from Henan Yuanzheng Special Gas Co. Ltd.

CO2-induced phase engineering

50.0 mg of pure CaSnO3 powder was directly transferred into the supercritical CO2 apparatus, composed mainly of a stainless-steel autoclave with a heating jacket and a temperature controller. The autoclave was heated to the designated temperature (100 °C), and then CO2 was charged into the reactor to the desired pressure (14/16/18 MPa) and maintained for 4 h under continual stirring. After the reactor had cooled, CO2 was slowly released and the solid product CaSnO3 was collected. Finally, the samples were dried in an oven at 80 °C.

Characterization

X-ray diffraction (XRD) patterns were collected on a Rigaku SmartLab SE diffractometer (Japan) using Cu K radiation. Transmission electron microscope (TEM) images were recorded on a JEOL JEM-F200 (Japan) at an acceleration voltage of 200 kV. Scanning electron microscope (SEM) and energy dispersive spectrometer (EDS) images were recorded on a ZEISS Sigma 360 (Germany). Raman measurements were performed using a Horiba LabRAM HR Evolution (Japan) with a laser wavelength of 532 nm. X-ray photoelectron spectroscopy (XPS) was performed using a Thermo Scientific K-Alpha system (America). The electron paramagnetic resonances (EPR) were obtained using an electron paramagnetic resonance spectrometer (EMX-9.5/12). Vibrating sample magnetometry (VSM) was performed using the LakeShore 7404 system (America). Field cooling–zero magnetic field cooling (FC–ZFC) curves were obtained using the Physical Characterisation Measurement System (China Quantum Design, PPMS-9).

Conclusions

In summary, SC CO2 treatment successfully alters the internal chemical environment of CaSnO3, enabling room-temperature ferromagnetism without the addition of magnetic elements. Experimental studies have demonstrated the critical synergistic role of SC CO2-induced defect engineering and strain effects in ferromagnetic manipulation. Specifically, EPR and XPS results confirm the existence of oxygen vacancies and Sn3+, directly showing that SC CO2 treatment introduces controllable lattice-scale defects in CaSnO3, thereby releasing residual force within distorted SnO6 octahedra. This stress relaxation suppresses octahedral distortion and drives a structural phase transition from orthorhombic to cubic, as unambiguously validated by XRD, TEM and SAED characterizations. The symmetry enhancement is accompanied by anisotropic lattice expansion and tensile strain, which lowers the oxygen vacancy formation energy, consequently disrupting the intrinsic antiferromagnetic order and promoting ferromagnetic coupling to achieve room-temperature ferromagnetic ordering. Beyond proposing a novel non-stoichiometric doping strategy for magnetic modulation, which is a universally applicable methodology that advances the design of room-temperature magnetic functional materials, this work elucidates a defect-strain synergy mechanism capable of tuning material magnetic order, offering a transformative pathway for developing novel functional materials for next-generation spintronic devices.

Author contributions

L. Xu conducted experiments, collected data and wrote the original draft. Q. Xu and B. Han designed the project and funded it, and also conducted writing – review and editing.

Conflicts of interest

There are no conflicts to declare.

Data availability

Data for this article are available from the corresponding authors upon reasonable request.

The data supporting this article have been included as part of the supplementary information (SI). Supplementary information: supplementary structural and microstructural characterization data, including refined XRD lattice parameters, Rietveld-refined patterns, TEM and SEM images, EDS elemental analysis, total spectrum of XPS characterization, and FC-ZFC magnetization curves. See DOI: https://doi.org/10.1039/d5sc04873g.

Acknowledgements

We are grateful to the National Natural Science Foundation of China (no. 51173170, 21703207, 21773216), the joint project from the Henan-Provincial and the China-National Natural Science Foundation (project no. U2004208), and the Central Plains Science and Technology Innovation Leading Talent Project (project no. 234200510008).

Notes and references

  1. C. C. Mayorga-Martinez, J. Zelenka, K. Klima, P. Mayorga-Burrezo, L. Hoang, T. Ruml and M. Pumera, ACS Nano, 2022, 16, 8694–8703 CrossRef CAS PubMed.
  2. J. Lu, J. He, G. Kang, B. Lan, J. Liu and M. Zhu, Tribol. Int., 2025, 202, 110305 CrossRef CAS.
  3. J. Chu, Y. Wang, X. Wang, K. Hu, G. Rao, C. Gong, C. Wu, H. Hong, X. Wang, K. Liu, C. Gao and J. Xiong, Adv. Mater., 2020, 33, 2004469 CrossRef PubMed.
  4. Z. Zhang, R. Sun and Z. Wang, ACS Nano, 2025, 19, 187–228 CrossRef CAS PubMed.
  5. H. M. Mohamed, I. A. Mabrouk, S. M. Yakout, Y. K. Abdel-Monem and W. Sharmoukh, Ceram. Int., 2025, 51, 1682–1689 CrossRef CAS.
  6. W. Sharmoukh, T. A. Hameed and S. M. Yakout, J. Alloys Compd., 2022, 925, 166702 CrossRef CAS.
  7. W. Sharmoukh and S. M. Yakout, J. Alloys Compd., 2024, 993, 174664 CrossRef CAS.
  8. C. Sun, J. A. Alonso and J. Bian, Adv. Energy Mater., 2021, 11, 2000459 CrossRef CAS.
  9. J. Yan, S. Zhang, Q. Wei, S. Cao, J. Zhao, B. Zou and R. Zeng, Adv. Opt. Mater., 2022, 10, 2101406 CrossRef CAS.
  10. A. G. Ricciardulli, S. Yang, J. H. Smet and M. Saliba, Nat. Mater., 2021, 20, 1325–1336 CrossRef CAS PubMed.
  11. J. Wang, C. Zhang, H. Liu, R. McLaughlin, Y. Zhai, S. R. Vardeny, X. Liu, S. McGill, D. Semenov, H. Guo, R. Tsuchikawa, V. V. Deshpande, D. Sun and Z. V. Vardeny, Nat. Commun., 2019, 10, 129 CrossRef PubMed.
  12. C. Zheng and Q. Liu, RSC Adv., 2019, 9, 33596–33601 RSC.
  13. Q. Liu, F. Jin, B. Li and L. Geng, J. Alloys Compd., 2017, 717, 55–61 CrossRef CAS.
  14. Y. Wang, W. Lei, S. Wu, F. Niu, Q. He, Y. Shen and F. Li, Ceram. Int., 2023, 49, 14426–14431 CrossRef CAS.
  15. S. Wi, M. Jeong, K. Lee and Y. Lee, Adv. Sci., 2024, 11, 2402848 CrossRef CAS PubMed.
  16. A. Manoharan, M. Munusamy, A. Pradeep, S. Sellaiyan, S. Hussain and S. Krishnan, Appl. Phys. A, 2020, 126, 874 CrossRef CAS.
  17. S. Sumithra and N. Victor Jaya, Mater. Res. Innov., 2019, 23, 375–384 CrossRef.
  18. A. A. Bhat and R. Tomar, J. Alloys Compd., 2021, 876, 160043 CrossRef CAS.
  19. Y. Ouyang, B. Gao, Y. Tang, L. Li and Q. Xu, Chem. Sci., 2025, 16, 1336–1343 RSC.
  20. D. Fu, M. Tan, Y. Liang and Q. Xu, Small, 2025, 2411243,  DOI:10.1002/smll.202411243.
  21. W. Liu, Q. Xu, W. Cui, C. Zhu and Y. Qi, Angew. Chem., Int. Ed., 2017, 56, 1600–1604 CrossRef CAS PubMed.
  22. Y. Zhou, Q. Xu, T. Ge, X. Zheng, L. Zhang and P. Yan, Angew. Chem., Int. Ed., 2020, 59, 3322–3328 CrossRef CAS PubMed.
  23. B. Gao, S. Xu and Q. Xu, Angew. Chem., Int. Ed., 2022, 61, e202117084 CrossRef CAS PubMed.
  24. T. Ge, W. Cui and Q. Xu, Angew. Chem., Int. Ed., 2023, 62, e202300446 CrossRef CAS PubMed.
  25. Y. Zhou, P. Yan, J. Jia, S. Zhang, X. Zheng, L. Zhang, B. Zhang, J. Chen, W. Hao, G. Chen, Q. Xu and B. Han, J. Mater. Chem. A, 2020, 8, 13320–13327 RSC.
  26. L. Du, B. Gao, S. Xu and Q. Xu, Nat. Commun., 2023, 14, 2278 CrossRef CAS PubMed.
  27. Ž. Knez, E. Markočič, M. Leitgeb, M. Primožič, M. Knez Hrnčič and M. Škerget, Energy, 2014, 77, 235–243 CrossRef.
  28. H. Shaili, E. Salmani, M. Beraich, M. Zidane, M. Taibi, M. Rouchdi, H. Ez-Zahraouy, N. Hassanain and A. Mzerd, ACS Omega, 2021, 6, 32537–32547 CrossRef CAS PubMed.
  29. H. P. Rooksby, Nature, 1945, 155, 484 CrossRef CAS.
  30. J. Gao, H. Zhang, L. Xie, J. Yu, G. Wang, Y. Gu, H. He and J. Bai, J. Eur. Ceram. Soc., 2017, 37, 985–994 CrossRef CAS.
  31. H. L. Zheng, Z. C. Zhang, J. G. Zhou, S. S. Yang and J. Zhao, Appl. Phys. A, 2012, 108, 465–473 CrossRef CAS.
  32. J. Kung, Y.-J. Lin and C.-M. Lin, J. Chem. Phys., 2011, 135, 224507 CrossRef PubMed.
  33. M. D. Williams, D. S. Bradshaw and D. L. Andrews, J. Chem. Phys., 2016, 145, 184301 CrossRef PubMed.
  34. C. Zhang, R. Lu, C. Liu, L. Yuan, J. Wang, Y. Zhao and C. Yu, Adv. Funct. Mater., 2021, 31, 2100099 CrossRef CAS.
  35. J. Liu, Y. Xiang, Y. Chen, H. Zhang, B. Ye, L. Ren, W. Tan, A. Kappler and J. Hou, Environ. Sci. Technol., 2023, 57, 12453–12464 CrossRef CAS PubMed.
  36. X. Song, H. Ma, Y. Li, Y. Duan, W. Wang and F. Dong, Appl. Catal., B, 2025, 365, 124955 CrossRef CAS.
  37. A. Manoharan, B. Suresh, k. Ajay, M. Munisamy and S. Krishnan, Solid State Commun., 2023, 360, 115033 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.