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

Impact of Bi3+ and Cu2+ doping on the optical and electronic properties of CsPbBr3 for photocatalytic toluene oxidation

Marija Knezevic a, Thi-Hieu Hoanga, Vien Duong Quacha, Alba Garzón Manjónc, David Llorens Rauretc, Marie Erarda, Audrey Gayrala, Mireille Benoita, David Berardanb, Jordi Arbiolcd, Christophe Colbeau-Justina and Mohamed Nawfal Ghazzal*a
aUniversité Paris-Saclay, UMR 8000 CNRS, Institut de Chimie Physique, 91405, Orsay, France. E-mail: mohamed-nawfal.ghazzal@universite-paris-saclay.fr
bUniversité Paris-Saclay, UMR 8182 CNRS, Institut de chimie moleculaire, F-91405, Orsay, France
cCatalan Institute of Nanoscience and Nanotechnology (ICN2), CSIC and BIST, Campus UAB, Bellaterra, 08193 Barcelona, Catalonia, Spain
dICREA, Pg. Lluís Companys 23, 08010 Barcelona, Catalonia, Spain

Received 8th June 2025 , Accepted 25th July 2025

First published on 28th July 2025


Abstract

Doping of metal halide perovskites (MHPs) offers an opportunity to introduce trap states that are reported to be beneficial for the charge carrier dynamics. Herein, we investigated the electronic structure–photocatalytic relationship of CsPbBr3 doped with Bi3+ and Cu2+ cations. Bi doping necessitates a reduction in the hot-injection temperature to avoid the formation of Bi nanoparticles, leading to the formation of two-dimensional CsPbBr3 with a variable aspect ratio. Bi doping causes no significant changes in the band gap energy but leads to the gradual appearance of a band-tail in the absorbance spectrum and a broad-band emission peak covering a large wavelength range. A photoluminescent quenching is observed as the ratio of Bi3+ increases, indicating shallow trap generation. Cu2+ doping does not affect the morphology of the 3D crystalline CsPbBr3 and shows a photoluminescent blueshift, which translates to a change in the bandgap energy. The CsPbBr3 doping, with an adequate amount of Cu, provides the best photocatalytic activity, surpassing Bi-doped CsPbBr3. Time-resolved photoluminescence analysis indicated that Bi3+ introduces a shallow trap that captures electrons while leaving the holes free, leading to an increase in the charge carrier lifetime, thus favoring charge carrier separation. Unlike bismuth, Cu2+ affects both the conduction and valence band positions, making the oxidation reaction more favorable.


Introduction

Metal halide perovskites (MHPs) have renewed the interest in solar energy conversion since they were included in dye-sensitized solar cells and other solar cell technologies.1,2 MHPs have versatile properties that offer plenty of possibilities to control their optoelectronic properties and light-harvesting capabilities.3 The most commonly reported is the partial anion substitution (Br1−x/Clx or Br1−x/Ix), which is an excellent way to tune the optical and electronic bandgaps of perovskites, leading to the fine control of photoluminescence, charge-carrier transfer, and photocatalytic properties.4 Partial halide substitution leads to an upward shift in the valence band maximum (VBM) and a downward shift in the conduction band minimum (CBM) as both have a contribution from the halide antibonding orbitals.5,6 Cation substitution has been envisaged mainly to diminish the toxicity of MHPs due to the presence of Pb in their structure.7 Partial replacement of Pb2+ cations in CsPbBr3 QDs with Sn2+, Cd2+, and Zn2+ was achieved by post-synthetic cation exchange.8,9 A partial substitution of Pb leads to either PL enhancement or quenching, depending on the oxidation potential of the metallic cation and whether or not the trap states are introduced.7,10,11 In addition, the complete replacement of lead through single and bimetallic substitution (Ag, Bi, Sn, and Sb) resulted in the formation of 2D double perovskites (A2MX6, A3M2X6, and A2M1M2X6).12–14 On the contrary, divalent and trivalent cations, such as Mn, Ni, Cu, Al, Ce, and Bi, could affect structural, optical, and electronic properties while retaining either the cubic or orthorhombic crystal structure of MHPs.10,11,15–17 The trivalent Bi3+ and Pb2+ have very close ionic radii and are isoelectronic (6s2). Bi3+ was employed to substitute Pb2+ and reduce the toxicity of 2D perovskite structures, which have shown enhanced photoactivity, e.g., 0D bismuth-based halide double perovskites such as Cs3Bi2I9,18 Cs3Bi2X9 embedded in mesoporous silica (Cs3Bi2X9-SBA-15),19 and Cs3Bi2X9 coupled with graphitic carbon nitride (Cs3Bi2X9-g-C3N4).20 Moreover, 2D Cs3Bi2X9 and Cs3Bi2X9-g-C3N4-Pt demonstrated interesting activity for H2 production from pure ethanol and 10% v/v aqueous triethanolamine, respectively.21,22 From another point of view, Bi-doped hybrid or inorganic perovskites have demonstrated promising optical and electronic properties.23–25 For example, doping CsPbBr3 perovskite nanocrystals with Bi3+ ions led to photoluminescence quantum yield (PLQY) quenching and a potential change in the bandgap energy.16 However, this change in bandgap energy has been questioned. In particular, the combination of spectroscopic ellipsometry and transmission spectroscopy suggested that the change in the absorption onset does not rationalize the change in the optical bandgap. Further direct measurement of the bandgap of Bi-doped CsPbBr3 using angle-resolved photoelectron spectroscopy revealed no change in the valence band energy.26 Upon Bi doping, the photoluminescence (PL) quenching was instead confirmed and associated with the formation of trap states that enhance electron trapping and suppress radiative recombination.26 In the same context, copper, with a smaller ionic radius (73 pm) compared to Pb2+ (119 pm), is another potential dopant for MHPs. Doping mixed-halide perovskites with divalent Cu2+ ions was reported to increase lattice binding energy and improve the short-range order of the lattice, which was identified as a factor improving both the thermal stability and PLQYs.7 The crystal lattice contraction through partial substitution of Pb2+ with Cu2+ and a reduction in surface energy in two-dimensional CsPbBr3 perovskite nanoplatelets after Cu doping lead to long-term stability and photostability under ambient conditions. Although Cu-doped all-inorganic perovskites have primarily been studied for their photoluminescence properties, there is limited research on their photocatalytic properties. To the best of our knowledge, only the photocatalytic activity of Cu/CsPbBr3–Cs4PbBr6 nanocrystals (NCs) has been studied.27

Herein, we investigated the effect of CsPbBr3 doping with bi- and tri-valent cations on its electronic and photocatalytic properties. This choice was motivated by the fact that Bi and Cu doping affect charge carriers differently, which is a determining parameter expected to modify the photocatalytic activity. Indeed, Bi has been reported to quench the PL, whereas Cu doping is more likely to enhance it. Thus, chemical, morphological, structural, optical, and electronic properties were investigated after the doping. Interestingly, varying the valence state of the dopants led to different effects, imparting each perovskite with specific properties. For instance, a drastic change in the morphology and the optical response was observed with Bi doping, whereas the perovskite maintained its morphology once doped with Cu. The charge carrier lifetime was determined using time-resolved laser scanning time-correlated single photon counting (TCSPC) microscopy, suggesting that both Cu and Bi doping could introduce shallow defects, thus affecting the photocatalytic efficiency in toluene oxidation.

Experimental section

Materials

All chemicals were used as received, without further purification or modification. Cesium carbonate (Cs2CO3, 99.5%), lead bromide (PbBr2, 99%), and oleylamine (OAm, 80%–90%) were purchased from Acros Organics. Oleic acid (OA) was obtained from Riedel de Haën. Toluene (99.8%), tetrabutylammonium hexafluorophosphate ((Bu4N)(PF6), ≥98%), dichloromethane (DCM, ≥99.8%), 1-octadecene (ODE, 90%), n-hexane (95%), and acetonitrile (CH3CN, ≥99.9%, for HPLC) were purchased from Sigma-Aldrich.

Synthesis of phase-pure CsPbBr3

CsPbBr3 nanocrystals were synthesized following a previously reported procedure.28,29 Briefly, Cs-oleate was prepared by mixing pre-dried Cs2CO3 (0.163 g), OA (1.58 mL), and ODE (3.42 mL) at 120 °C under an argon flow for 1 h with vigorous stirring at 500 rpm. In order to completely dissolve Cs2CO3, the temperature was increased up to 130 °C at a rate of 1 °C min−1. Cs-oleate must be preheated to 120 °C before use in the synthesis of CsPbBr3 to avoid precipitation at room temperature. Simultaneously, PbBr2 (0.21 g), OA (1.5 mL), OAm (1.5 mL), and ODE (5 mL) were mixed and heated at 120 °C under an argon flow for 1 h. After complete dissolution of PbBr2, the temperature was raised to 200 °C. The previously prepared and preheated solution of Cs-oleate was then injected (0.75 mL). The reaction was immediately quenched by immersing the flask in an ice/water bath. CsPbBr3 was dispersed in toluene and centrifuged at 7500 rpm for 20 min. The supernatant was discarded, and CsPbBr3 NCs were washed three times with n-hexane (total 10 mL) and centrifuged again at 7500 rpm for 10 min. After the removal of aggregates, the elaborated CsPbBr3 was dispersed in n-hexane.

Synthesis of Bi-doped CsPbBr3 nanocrystals

The bismuth-doped CsPbBr3 MHPs were synthesized by adapting the synthetic procedure used for pristine CsPbBr3. Bi doping was introduced by partially substituting Pb2+ with Bi3+ to yield Cs2Pb1−xBixBr3 compositions, where x = 0.01, 0.02, 0.05, and 0.10, corresponding to 1, 2, 5, and 10 mol% of Bi relative to the total B-site cations (Pb + Bi). The amount of PbBr2 was fixed at 0.21 g, and the mass of BiBr3 was varied accordingly to achieve the desired stoichiometric ratios. The variable mass ratio of BiBr3 and PbBr2 was mixed with OA (3 mL), OAm (1.5 mL), and ODE (10 mL), and heated at 120 °C, stirring at 500 rpm under argon flow for 1 h. After the complete dissolution of BiBr3 and PbBr2, the temperature was increased to 160 °C at a rate of 1 °C min−1. The previously prepared and preheated solution of Cs-oleate was injected (0.75 mL). The reaction was immediately quenched by immersing the flask in an ice/water bath. The purification process using hexane and toluene for the bismuth-doped CsPbBr3 was identical to that of the pristine CsPbBr3. The synthesis of bismuth-doped CsPbBr3 was also carried out at 200 °C, using equivalent amounts of OA (3 mL), OAm (3 mL), and ODE (10 mL).

Synthesis of Cu-doped CsPbBr3 nanocrystals

The copper-doped CsPbBr3 was synthesized by mixing a variable molar ratio (Pb[thin space (1/6-em)]:[thin space (1/6-em)]Cu = 1[thin space (1/6-em)]:[thin space (1/6-em)]0.25, 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5, 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5) of CuBr2 and PbBr2 (0.21 g) with OA (1.5 mL), OAm (1.5 mL), and ODE (10 mL), and heating the mixture at 120 °C under argon flow for 1 h. After the dissolution of CuBr2 and PbBr2, the temperature was increased to 200 °C. The previously prepared and preheated solution of Cs-oleate (0.75 mL) was injected. The reaction was immediately quenched by immersing the flask in an ice/water bath. The purification process with hexane and toluene was identical to that of pristine CsPbBr3.

Characterization techniques

Diffuse reflectance spectra (DRS) were recorded using a UV-vis-NIR Cary 5000 spectrophotometer (Agilent Technologies) equipped with an integrating sphere for diffuse and total reflection measurements. The maximum reflectance was set to 100% using BaSO4 as a reference in the 200 to 1100 nm wavelength range.

Photoluminescence (PL) emission spectra were recorded using a Fluorolog-3 spectrofluorometer (HORIBA) equipped with a solid-state sample holder. All samples were excited at 400 nm, and emission spectra were monitored over the 450 to 700 nm range.

X-ray photoelectron spectroscopy (XPS). Measurements were performed on a K-Alpha spectrometer from ThermoFisher, equipped with a monochromated X-ray Source (Al Kα, 1486.6 eV) with a spot size of 400 μm. The hemispherical analyzer was operated in constant analyser energy (CAE) mode, with a pass energy of 200 eV and a step size of 1 eV for the acquisition of survey spectra, whereas for the acquisition of narrow scans, a pass energy of 50 eV and 100 eV and a step size of 0.1 eV were used. The 50 eV pass energy was utilized for core levels requiring high spectral resolution to accurately resolve chemical shifts and fine structure, as seen with Bi 4f and Cu 2p. In contrast, we applied a pass energy of 100 eV for core levels characterized by lower signal intensity or overlapping peaks. This approach allowed us to enhance signal strength while still maintaining a reasonable level of resolution.

The charge build-up was neutralized by a “dual-beam” flood gun. The obtained spectra were treated using CasaXPS software. The binding energies were calibrated against the C 1s binding energy set at 284.8 eV.

Powder X-ray diffraction (PXRD) patterns were recorded using a room-temperature powder X-ray diffractometer. Patterns were recorded by a Panalytical X'Pert diffractometer equipped with a Ge(111) incident monochromator (Cu-Kα radiation) and an X'celerator detector.

Transmission electron microscopy (TEM). The TEM and HRTEM images were obtained using a JEOL JEM 2100Plus operating at 200 kV and a Tecnai F20 field emission gun (FEG) microscope at 200 kV. The size of synthesized nanostructures was measured using Gatan Digital Micrograph software. All the crystal lattices in the micrographs of interest were enhanced using a frequency filter in the reciprocal space.
Electron energy loss spectroscopy (EELS). Compositional analysis was conducted using a double-corrected and monochromated Thermo Fisher Spectra 300 microscope operated at 300 kV. For elemental mapping and quantification studies, the L edges of Br and M4,5 edges of Cs and Pb were used. To reduce noise in the EELS spectra, principal component analysis (PCA) was conducted, followed by reconstruction using a sufficient number of components to prevent data loss. EELS data analysis was conducted using Gatan Digital Micrograph software.

Time-resolved photoluminescence lifetime measurements were carried out with a homemade setup. All samples were excited at 370 nm (80 ps FWHM, 10 MHz, PicoQuant). The instrument response function was determined by the excitation of colloidal silica (SiO2, Ludox) at 370 nm. The emitted luminescence was selected using a monochromator (ARC Spectra drive) at 517 ± 5 nm and detected by an MCP-PMT (Hamamatsu) connected to a PicoHarp 300 TCSPC module (PicoQuant). Lifetime measurements were analyzed using the SymPhoTime software (PicoQuant).

Cyclic voltammetry (CV) measurements were carried out using (Bu4N)(PF6) 0.1 M in dichloromethane (DCM) as the electrolyte. The formal potential of the non-aqueous reference electrode Ag|Ag+ was calibrated by adding 2 mM ferrocene/ferrocenium (Fc/Fc+) as an internal standard at the end of the measurements. All measurements were carried out using a PGSTAT101 Metrohm Autolab potentiostat.

Photocatalytic oxidation of toluene was performed by dispersing 3 mg of Bi-doped and Cu-doped CsPbBr3 NCs in 3 mL of toluene in a quartz reactor. The obtained dispersion was purged with a continuous flow of oxygen gas for 20 min. The samples were irradiated using an Oriel 300 W xenon lamp equipped with an infrared water filter for 3 h under stirring. The aliquots were taken every 30 min and analyzed using high-performance liquid chromatography (HPLC).

Results and discussion

We prepared undoped and doped CsPbBr3 by a hot-injection method. The obtained Cu-CsPbBr3 nanoparticles demonstrated a similar yellow color with varying PbBr2/CuBr2 ratios in the crude solution, i.e., the molar ratio of Pb/Cu was set to 0.25, 0.5, 1.0, and 1.5. Cu doping of MHPs required a significant amount of Cu(II) precursor due to the higher formation energy of Cu–Br compared to Pb–Br, e.g., −14.2 eV for CsCuBr3, −6.82 eV for CsPb0.93Cu0.07Br3, and −6.56 eV for CsPbBr3.7 Bismuth-doped cesium-lead bromide (hereafter Bi-CsPbBr3) nanoparticles were first synthesized at 200 °C by a hot-injection method. Higher temperatures were found to induce the reduction of Bi precursors and the formation of bismuth nanoparticles alongside perovskite nanocrystals (Fig. S1). Bismuth reduction occurred at high temperatures in a cesium-free solution in the presence of reductive agents and capping agents essential for controlling the size of MHPs, such as OA and OAm.30–32 Thus, the synthesis temperature was decreased to 160 °C, and the ratio of OA to OAm was adjusted (1[thin space (1/6-em)]:[thin space (1/6-em)]2) to avoid thermal reduction of bismuth.33,34 The color of the obtained Bi-CsPbBr3 nanoparticles changed from yellow (CsPbBr3) to dark orange (CsPb0.90Bi0.10Br3 – all compositions given in the article are nominal compositions), as the amount of Bi in the crude solution increased (1%, 2%, 5%, and 10%) (Fig. 1).
image file: d5nr02442k-f1.tif
Fig. 1 Photographs of Bi- and Cu-doped CsPbBr3 powders obtained after synthesis.

The absorbance spectra show that the excitonic band remains unaffected upon doping (Fig. 2a). Cu doping results in a blue-shifted emission as the molar ratio of Cu increases, compared with the pristine CsPbBr3 (Fig. 2b), which is in agreement with previous reports.7,35 The Kubelka–Munk calculation indicates no significant change in the optical bandgap (Fig. S2), which contrasts with the PL shift to a shorter wavelength (Fig. 2b). Cyclic voltammetry measurements were carried out for all samples, demonstrating changes in the potential energies of the VBM and CBM (Fig. 2c). The bandgap energy of CsPbBr3 increases from 2.31 eV to 2.50 eV upon Cu doping (Fig. 2d), aligning with the PL results. We attributed this discrepancy to the additional states introduced as shallow trap states in the CBM, which could not be observed in UV-Vis spectroscopy.


image file: d5nr02442k-f2.tif
Fig. 2 (a) UV-visible absorption spectra and (b) emission spectra of Cu-doped CsPbBr3 nanoparticles. (c) Cyclic voltammograms of Cu-doped CsPbBr3 recorded at 50 mV s−1 in a 100 mM solution of (Bu4N)(PF6) in DCM, showing oxidation (A1) and reduction (C1) peaks. (d) Energy band edge diagrams of Cu-CsPbBr3 MHPs deduced from (c). Electrochemical bandgap values (Eqpg, black) were compared with optical ones (Eoptg, pale pink).

Before discussing the Bi doping, we recorded the PL spectra of undoped CsPbBr3 synthesized at different temperatures (160 °C and 200 °C) and at 160 °C with an excess OA to enhance the solubility of the bismuth precursor (Fig. 3a). A blue shift of 11 nm in the PL maximum was observed when the synthesis temperature was lowered. An additional shift of 6 nm was noted when excess OA was used. Interestingly, the PL spectrum of the MHP synthesized with excess OA exhibited three additional emission peaks with lower intensity. The lowest-energy emission peak at 460 nm could be attributed to biexciton (BX), while the highest-energy exciton at 491 nm could be assigned to the active bright band (DX).36,37 This PL emission behavior is typically observed in 2D perovskites.36,37 The PL spectra of Bi-doped CsPbBr3 displayed a broad emission band covering a large emission range between 420 and 540 nm (Fig. 3b). The band edge emission was found to be composite, consisting of multi-emission peaks centered at 430, 465, 480, 490, and 520 nm, with the intensity decreasing as the Bi content increased. Additional emission peaks appear compared to undoped CsPbBr3, suggesting a supplementary excitonic phenomenon that could not be assigned at this stage. As the doping ratio increases from 1% to 2%, the peaks at 490 and 520 nm are quenched. At a higher doping ratio (10%), all peaks disappeared, and a new band edge emission appeared at 435 nm. Upon Bi doping, the excitonic absorption band is observed around 540 nm, similar to that of undoped CsPbBr3. Further increase in the Bi ratio leads to the gradual appearance of a band-tail in the absorbance spectrum, as shown in Fig. 3c. We evaluated the bandgap energy using the Kubelka–Munk function, and all Bi-CsPbBr3 samples demonstrated bandgap energies in the range of 2.2–2.3 eV (Fig. S2b). The conduction and valence band positions were determined using cyclic voltammetry measurements (Fig. S3). The obtained values, shown in Fig. 3d, are slightly higher than values obtained using the Kubelka–Munk calculation, falling in the range of 2.40–2.46 eV. This discrepancy is due to the presence of multi-emission peaks and the formation of trap states, which may account for the band-tail observed in the 550–700 nm range of the absorbance spectrum.38,39 Note that the optical bandgap describes the transition of the excitonic band of the electron–hole pair (eqn (1)):

 
Eopt = Eg + Eb (1)
where Eb represents the exciton binding energy and Eg represents the electronic bandgap.40 The exciton binding energy represents the coulombic interaction between electrons and holes and is typically observed as a smaller bandgap. The difference between the optical and electronic bandgap corresponds to the excitonic states located below the bandgap minimum, indicating that Bi doping introduces shallow trap states below the CBM. It is worth noting that Bi-doped CsPbBr3 demonstrated an upward shift of the conduction bandgap minimum (CBM), while the valence band maximum (VBM) remains at an energy level similar to that of pristine CsPbBr3 synthesized at 160 °C. Furthermore, the average bandgap energy estimated from the CV measurements is 0.3 eV higher than that of pristine CsPbBr3, consistent with previous studies.26


image file: d5nr02442k-f3.tif
Fig. 3 (a) Emission spectra of CsPbBr3 synthesized at 160 °C, 200 °C, and 160 °C with modified OA[thin space (1/6-em)]:[thin space (1/6-em)]OAm ratio. (b) Photoluminescence spectra of variable samples of Bi-doped CsPbBr3. (c) Absorption spectra of variable Bi-doping CsPbBr3 obtained at 160 °C. (d) Electrochemical bandgap values (Eqpg, black) were compared with optical ones (Eoptg, pale pink).

X-ray diffraction (XRD) patterns of each CsPbBr3 powder obtained using different doping amounts of Cu showed no extra peaks (Fig. 4a) compared with the undoped CsPbBr3 nanocrystal (Fig. S4), indicating an orthorhombic single-phase material that retains the structural framework of pure CsPbBr3 crystals, which were further confirmed using transmission electron microscopy (TEM) and fast Fourier transform (FFT) analysis (Fig. S5 and S6). The morphology of Cu-doped CsPbBr3 was observed by TEM, and the results are shown in Fig. 4b and Fig. S7a–c. TEM images revealed cubic morphologies of Cu-doped CsPbBr3 nanoparticles, with average sizes of 37, 20, and 21 nm when the ratio between Pb[thin space (1/6-em)]:[thin space (1/6-em)]Cu was 1[thin space (1/6-em)]:[thin space (1/6-em)]0.25, 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5 and 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5, respectively (Fig. S7). The CuBr2 treatment of CsPbBr3 at 200 °C did not affect the shape and size of Cu-CsPbBr3, except at the 1[thin space (1/6-em)]:[thin space (1/6-em)]0.25 doping level, where the particle size was twice that of pristine MHPs (∼20 nm). For the Pb[thin space (1/6-em)]:[thin space (1/6-em)]Cu = 1[thin space (1/6-em)]:[thin space (1/6-em)]1 sample, we could also assign orthorhombic CsPbBr3 (ICSD: 97851) NPs (Fig. S8a–e). The nanoparticles were oriented along the [101] or [210] zone axes (Fig. S8d and e). Similarly, at a higher doping ratio (1[thin space (1/6-em)]:[thin space (1/6-em)]1.5), the FFT analysis of the image revealed that the NPs adopted an orthorhombic CsPbBr3 structure. Despite Cu doping, we did not observe significant changes in the d-spacings and angles with respect to the reported CsPbBr3 crystal structure (Fig. 4b and Fig. S9), which is consistent with the XRD analysis, showing no evidence of any significant change in the lattice parameters of the orthorhombic structure. The Cu-doped CsPbBr3 nanoparticles maintained the same zone axis orientation. Moreover, energy-dispersive X-ray spectroscopy (EDS) mapping demonstrated a homogeneous distribution of Cs, Pb, and Br atoms (Fig. 4d). Electron energy loss spectroscopy (EELS), combined with principal component analysis (PCA), was applied to the image shown in Fig. 4c, to analyze elemental signal within the sample. Close-up spectra of individual elemental edges are shown in Fig. S10. Note that no Cu signal from the L3,2 edges was detected, which should appear at 931 and 951 eV. We can hypothesize that the intense signal from Cs might be suppressing the detection of the Cu signal.


image file: d5nr02442k-f4.tif
Fig. 4 (a) Powder XRD patterns of Cu-doped CsPbBr3 nanocrystals prepared from solutions containing variable amounts of the dopant precursor. (b) HR-TEM images of CsPbBr3[thin space (1/6-em)]:[thin space (1/6-em)]Cu = 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5 and its FFT analysis. (c) EDS elemental mappings of CsPbBr3[thin space (1/6-em)]:[thin space (1/6-em)]Cu = 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5. (d) EELS analysis of the selected region from (c). Note that the crystal lattices in the micrograph of the region of interest have been enhanced using a frequency filter in the reciprocal space.

X-ray diffraction (XRD) patterns indicate that the perovskite retains its orthorhombic structure upon treatment with a variable ratio of BiBr3 (Fig. 5a). The synthesis of Bi-doped MHP was performed at 160 °C. Thus, for comparison, TEM images of undoped CsPbBr3 are shown in Fig. S11. The images show well-dispersed nanocrystals with an average size of 16 nm. The morphology of CsPbBr3 shows the formation of CsPbBr3 nanosheets with a limited number of nanocubes.41 In contrast, we observe that the nanosheets are thinner than the nanocubes, with variable aspect ratios. Higher TEM magnification of Bi-doped CsPbBr3 indicates that the hot-injection synthesis at 160 °C yields nanocubes of variable sizes, ca. 10, 20, and 40 nm, with no formation of Bi nanoparticles. The shape of the MHP appears to change as the doping ratio increases, leading to the formation of 2D perovskite with variable aspect ratios (Fig. 5b, c and Fig. S12a–h). These 2D perovskites are mainly observed at lower doping ratios (less than 10%). The broad size distribution and the formation of the 2D perovskite rationalize the multiple-emission spectra observed in Fig. 3b. The 2D perovskite comprises varying aspect ratios and retains some 3D perovskite, leading to variable exciton relaxation behavior and translating variable photoluminescence peaks. A reduced dimension or thinner layer results in blue-shifted photoluminescence and a lower bandgap value. The FFT analysis of the image reveals that the NPs adopt an orthorhombic CsPbBr3 structure (ICSD: 97851), consistent with the XRD patterns (Fig. 5e). Nanoparticles were oriented along the [101] zone axis.


image file: d5nr02442k-f5.tif
Fig. 5 (a) Powder XRD patterns of Bi-doped CsPbBr3 nanocrystals prepared using solutions containing variable amounts of the dopant precursor. TEM images of CsPbBr3 doped with (b) 1%, (c) 5%, and (d) 10% of Bi obtained at 160 °C. (e) HR-TEM images of CsPbBr3 doped with 10% Bi and its FFT analysis. The crystal lattices in the micrograph of the region of interest have been enhanced using a frequency filter in the reciprocal space.

X-ray photoelectron spectroscopy (XPS) surface analysis was performed to demonstrate the presence of Bi and Cu in the doped perovskite. The survey and the high-resolution spectra of the samples indicate the presence of Cs, Pb, and Br in all doped and undoped CsPbBr3 (Fig. S13 and S14). High-resolution XPS spectra show the presence of oxygen in CsPbBr3, which could originate from OA binding to the surface, with a signal located at 532.1 eV (Fig. S14). The Cs 3d spectra of doped CsPbBr3 originate from 3d3/2 and 3d5/2 spin–orbit coupling, located at 724.5 and 738.4 eV, respectively (Fig. 6a). The Pb 4f peaks are split into 4f7/2 and 4f5/2, located at 138.2 and 143.1 eV, respectively (Fig. 6a). The high-resolution Br 3d spectra are centered at 68.1 and 68.3 eV (Fig. 6a). All peaks are similar to those obtained for Cu-doped perovskites and do not exhibit any significant shift compared to pristine CsPbBr3 (Fig. S14). XPS analysis confirms the presence of Bi3+ as the doping is performed. The Bi 4f5/2 component is distinguishable at a binding energy 164.3 eV, as shown in Fig. 6b. The Bi 4f region splits into two components, including Bi 4f7/2, which obviously overlaps with the Cs 4P3/2 component at 158.9 eV.11,42 The Cu-doped perovskites exhibit Cu 2p transitions, with two main contributions located at around 931.3 and 936.0 eV, as shown in Fig. 6c. These contributions are likely to translate into contributions that can be attributed to Cu2+ species in different coordination.43,44 The first peak is shifted to lower energy compared to that reported for tetracoordinated Cu. However, the peak centered at ca. 937 eV corresponds to isolated Cu2+ in octahedral coordination, consistent with the substitution of the Pb atom into the perovskite structure. The Cu 2p region overlaps with a peak at 928.1 eV, assigned to Cs MMNT.


image file: d5nr02442k-f6.tif
Fig. 6 High-resolution XPS spectra of (a) Cs, Pb, and Br in Bi-doped CsPbBr3, (b) Bi in Bi-doped CsPbBr3, and (c) Cu in Cu-doped CsPbBr3.

The photocatalytic activities of Bi- and Cu-doped CsPbBr3 NCs were evaluated for toluene oxidation under xenon lamp illumination for 3 h (Fig. 7). The photocatalytic activity showed a significant improvement in the oxidation of toluene due to Bi doping (Fig. 7a). We observed enhanced photocatalytic activity with Bi doping of CsPbBr3, irrespective of the ratio. The photocatalytic efficiency exhibited a maximum oxidation rate at 1% and 10% doping levels. The linear conversion trend of toluene to benzaldehyde during 3 h of illumination indicates the stability of the photocatalyst throughout the reaction. The photocatalytic oxidation rate increased from 171 ppm g−1 min−1 for undoped CsPbBr3 to 435 and 411 ppm g−1 min−1 for CsPb0.99Bi0.01Br3 and CsPb0.90Bi0.10Br3, respectively (Fig. 7b). We also assessed the photocatalytic properties of Cu-doped CsPbBr3 NCs. The photocatalytic toluene oxidation improved significantly compared to undoped CsPbBr3. The photocatalytic activity increased gradually, reaching an optimum at a Pb/Cu ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1, with a conversion of 80 × 103 ppm g−1 after 3 h of reaction. However, excessive doping (1[thin space (1/6-em)]:[thin space (1/6-em)]1.5) led to a drop in the benzaldehyde production, which was inferior to that of the undoped CsPbBr3 NCs. The rate of toluene photocatalytic oxidation reached 460 ppm g−1 min−1,1 closely comparable to that of Bi-doped CsPbBr3 (Fig. 7b).


image file: d5nr02442k-f7.tif
Fig. 7 Photocatalytic activity and toluene conversion rate under UV-Vis illumination for 3 hours over Bi-doped CsPbBr3 (a and b) and Cu-doped CsPbBr3 (c and d) photocatalysts.

The photocatalytic oxidation of toluene under UV-Vis illumination produces benzaldehyde as the main product, with benzylic alcohol as an intermediate. The key to successful toluene oxidation is pre-saturation with oxygen, which allows the generation of superoxide anion radicals (O2˙) in the presence of electrons (eqn (2)). The rate-determining step in toluene oxidation involves C(sp3)–H bond activation of adsorbed toluene on the photocatalyst surface, upon interaction with holes on the MHPs site (eqn (3)).19,45–48

 
O2 + e → O2˙ (2)
 
C6H5CH3 + h+ → C6H5CH2˙ (3)

The proposed mechanism of toluene oxidation is given by the following equations:

 
C6H5CH2˙ + O2 → C6H5CH2OO˙ (4)
 
C6H5CH2OO˙ + C6H5CH2˙ → C6H5CH2OH + C6H5CHO (5)
 
C6H5CH2˙ + O2˙ → C6H5CH2OO (6)
 
image file: d5nr02442k-t1.tif(7)

Photogenerated charge carriers play a major role in toluene oxidation. To investigate the effects of bismuth and copper on these carriers, time-resolved photoluminescence spectroscopy was performed. When CsPbBr3 was doped with bismuth, its photoluminescence was dramatically quenched. The average PL lifetime decreased from 17.5 ns in undoped CsPbBr3 (160 °C) to 1.3 ns (Fig. 8a and Table S1). On the other hand, undoped CsPbBr3 nanosheets prepared under the same conditions exhibited an even longer lifetime of 28.9 ns. Bismuth introduces shallow states just below the conduction band minimum, acting as electron-type trapping states.11 These additional states can trap electrons, leading to PL quenching. Copper doping gradually reduced the photoluminescence lifetime, but an excessive Cu content (Pb[thin space (1/6-em)]:[thin space (1/6-em)]Cu = 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5) led to an increase in lifetime, contrary to earlier reports that show enhancement at low doping and suppression at higher levels (Fig. 8b and Table S2).7 The PL lifetime decay profile was fitted using a multiexponential fit, and the average lifetime decreased to 1.4 ns at a Pb[thin space (1/6-em)]:[thin space (1/6-em)]Cu ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1. Interestingly, at Pb[thin space (1/6-em)]:[thin space (1/6-em)]Cu = 1[thin space (1/6-em)]:[thin space (1/6-em)]1.5, the lifetime increased to 6.2 ns, although this value remains much lower than that observed for undoped CsPbBr3.


image file: d5nr02442k-f8.tif
Fig. 8 TRPL lifetime decay of (a) Bi-doped CsPbBr3 and (b) Cu-doped CsPbBr3 upon laser excitation at 375 nm.

In summary, low bismuth doping generates shallow traps that capture electrons and facilitate the formation of oxygen radicals, thereby participating in charge carrier separation, which is reflected in the reduction of photoluminescence (PL) lifetime. Consequently, holes remain free for the activation of C(sp3)–H bonds, resulting in the generation of benzylic radicals. However, further doping with bismuth does not yield a beneficial effect on the photocatalytic oxidation of toluene, except at a higher bismuth doping ratio. At elevated doping levels, the effects are expected to be more pronounced due to the reduced PL lifetime, which leads to enhanced photocatalytic activity. Additionally, copper appears to act as an electron collector, although to a limited extent. Indeed, the quenching of PL is interpreted as a result of copper creating shallow traps that capture electrons, thus leaving free holes for the oxidation reaction. Nevertheless, an excessive loading of copper leads to photocatalytic deactivation, which may originate from the formation of deep traps within the band structure that function as recombination sites.

Conclusion

Herein, we achieved Bi and Cu doping of CsPbBr3 using the hot-injection method. We found that the effect of doping on the optical, electronic, morphological, and photocatalytic activity depends on the nature of the cation used. For Bi doping of CsPbBr3, only a small amount of BiBr3 is necessary to initiate the reaction. It is crucial to maintain the synthetic reaction temperature at 160 °C to avoid the Bi reduction and formation of metallic nanoparticles. A reduction in the temperature led to the formation of 2D perovskite nanosheets with varying aspect ratios. However, doping with copper requires a significantly higher amount of CuBr2 to initiate doping owing to the higher energy required to form the Cu–Br bond compared to the Pb–Br bond. We found that the bandgap energy increased with Bi and Cu doping, indicating the introduction of a trap state. The band-tail in the absorption spectrum of Bi-doped MHP suggests the generation of shallow trap states in the conduction band. Both metals indicate photoluminescence quenching, demonstrating charge transfer from MHP to either the Bi or Cu trap states. We found that the photocatalytic activity toward toluene oxidation improved upon doping. However, the mechanism behind this enhancement varies depending on the cation used. Copper doping exhibited remarkable photocatalytic activity and stability during 3 h reactions. The optimal photocatalytic activity was achieved at a nominal Cu ratio of 0.5 (Pb[thin space (1/6-em)]:[thin space (1/6-em)]Cu = 1[thin space (1/6-em)]:[thin space (1/6-em)]0.5), surpassing that of Bi-doped MHP. Unlike bismuth, which incorporates trap states below the conduction band minimum (CBM), copper influences both the conduction and valence band positions, making the oxidation reaction more favorable.

Author contributions

M. Knezevich and T.-H. Hoang carried out the experiments and investigation process, curated the data, wrote the initial draft, and reviewed and edited the final manuscript. V. D. Quach performed the PEC experiments. A. G. Manjón, D. Llorens Rauret, and J. Arbiol performed the electron microscopy analyses and discussed their results. M. Erard and A. Gayral performed the TRPL measurements and discussed the results. D. Berardan performed the XRD analysis and discussed the result. M. Benoit supervised the HPLC analysis and set up the analytical method. C. Colbeau-Justin supervised the study. M. N. Ghazzal conceptualized and conceived the idea, acquired the financing, supervised the work, contributed to the manuscript structure, and reviewed and edited the final manuscript. The manuscript was written with the help of all the co-authors.

Conflicts of interest

The authors declare that no conflict of interest, financial or otherwise, exists.

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Supplementary information such as TEM, Kubelka-Munk calculations, XRD, Cyclic voltammograms, EELS spectra, XPS and TRPL is available. See DOI: https://doi.org/10.1039/d5nr02442k.

Acknowledgements

MK acknowledges the French “Ministère de l'enseignement supérieur et de la recherche” (MESR) for the PhD Grant. MNG thanks the public grant overseen by the French National Research Agency (ANR) through the IngenCat project (ANR-20-CE43-0014) and the NEXTCCUS project as part of the ERANET-ACT3 call program for the financial support. The authors thank François Brisset for his valuable help in performing the TEM analysis. The authors thank Diana Dragoe for her valuable help in performing the XPS analysis. ICN2 is supported by the Severo Ochoa program from the Spanish MCIN/AEI (grant no.: CEX2021-001214-S) and is funded by the CERCA Programme from the Generalitat de Catalunya. ICN2 acknowledges funding from the Generalitat de Catalunya 2021SGR00457. The authors thank for the support from the project AMaDE (PID2023-149158OB-C43), funded by MCIN/AEI/10.13039/501100011033/ and by “ERDF A way of making Europe”, by the “European Union”. AGM acknowledges the RyC Programme by MCIN/AEI and Ayuda RYC2021-033479-I funded by MCIN/AEI/10.13039/501100011033 and by the European Union NextGenerationEU/PRTR. Part of the present work was performed in the framework of Universitat Autònoma de Barcelona Materials Science PhD program. ICN2 is a founding member of e-DREAM.49

References

  1. I. Chung, B. Lee, J. He, R. P. H. Chang and M. G. Kanatzidis, Nature, 2012, 485, 486–489 CrossRef CAS PubMed .
  2. F. Hao, C. C. Stoumpos, D. H. Cao, R. P. H. Chang and M. G. Kanatzidis, Nat. Photonics, 2014, 8, 489–494 CrossRef CAS .
  3. M. Knezevic, T.-H. Hoang, N. Rashid, M. Abdi-Jalebi, C. Colbeau-Justin and M. N. Ghazzal, Sci. China Mater., 2023, 66, 2545–2572 CrossRef CAS .
  4. V. K. Ravi, G. B. Markad and A. Nag, ACS Energy Lett., 2016, 1, 665–671 CrossRef CAS .
  5. M. V. Kovalenko, L. Protesescu and M. I. Bodnarchuk, Science, 2017, 358, 745–750 CrossRef CAS PubMed .
  6. M. Knezevic, V.-D. Quach, I. Lampre, M. Erard, P. Pernot, D. Berardan, C. Colbeau-Justin and M. N. Ghazzal, J. Mater. Chem. A, 2023, 11, 6226–6236 RSC .
  7. C. Bi, S. Wang, Q. Li, S. V. Kershaw, J. Tian and A. L. Rogach, J. Phys. Chem. Lett., 2019, 10, 943–952 CrossRef CAS PubMed .
  8. N. Mondal, A. De and A. Samanta, ACS Energy Lett., 2019, 4, 32–39 CrossRef CAS .
  9. W. Van Der Stam, J. J. Geuchies, T. Altantzis, K. H. W. Van Den Bos, J. D. Meeldijk, S. Van Aert, S. Bals, D. Vanmaekelbergh and C. De Mello Donega, J. Am. Chem. Soc., 2017, 139, 4087–4097 CrossRef CAS PubMed .
  10. A. Shapiro, M. W. Heindl, F. Horani, M.-H. Dahan, J. Tang, Y. Amouyal and E. Lifshitz, J. Phys. Chem. C, 2019, 123, 24979–24987 CrossRef CAS .
  11. R. Begum, M. R. Parida, A. L. Abdelhady, B. Murali, N. M. Alyami, G. H. Ahmed, M. N. Hedhili, O. M. Bakr and O. F. Mohammed, J. Am. Chem. Soc., 2017, 139, 731–737 CrossRef CAS PubMed .
  12. B. Yang and K. Han, Acc. Chem. Res., 2019, 52, 3188–3198 CrossRef CAS PubMed .
  13. E. M. Hutter, M. C. Gélvez-Rueda, D. Bartesaghi, F. C. Grozema and T. J. Savenije, ACS Omega, 2018, 3, 11655–11662 CrossRef CAS PubMed .
  14. D. Bartesaghi, A. H. Slavney, M. C. Gélvez-Rueda, B. A. Connor, F. C. Grozema, H. I. Karunadasa and T. J. Savenije, J. Phys. Chem. C, 2018, 122, 4809–4816 CrossRef CAS PubMed .
  15. Y.-W. Liu, S.-H. Guo, S.-Q. You, C.-Y. Sun, X.-L. Wang, L. Zhao and Z.-M. Su, Nanotechnology, 2020, 31, 215605 CrossRef CAS PubMed .
  16. J. Yin, G. H. Ahmed, O. M. Bakr, J.-L. Brédas and O. F. Mohammed, ACS Energy Lett., 2019, 4, 789–795 CrossRef CAS .
  17. M. Liu, G. Zhong, Y. Yin, J. Miao, K. Li, C. Wang, X. Xu, C. Shen and H. Meng, Adv. Sci., 2017, 4, 1700335 CrossRef PubMed .
  18. B.-M. Bresolin, C. Günnemann, D. W. Bahnemann and M. Sillanpää, Nanomaterials, 2020, 10, 763 CrossRef CAS PubMed .
  19. Y. Dai, C. Poidevin, C. Ochoa-Hernández, A. A. Auer and H. Tüysüz, Angew. Chem., 2020, 132, 5837–5845 CrossRef .
  20. M. Miodyńska, A. Mikolajczyk, P. Mazierski, T. Klimczuk, W. Lisowski, G. Trykowski and A. Zaleska-Medynska, Appl. Surf. Sci., 2022, 588, 152921 CrossRef .
  21. Y. Ji, M. She, X. Bai, E. Liu, W. Xue, Z. Zhang, K. Wan, P. Liu, S. Zhang and J. Li, Adv. Funct. Mater., 2022, 32, 2201721 CrossRef CAS .
  22. L. Romani, A. Speltini, C. N. Dibenedetto, A. Listorti, F. Ambrosio, E. Mosconi, A. Simbula, M. Saba, A. Profumo, P. Quadrelli, F. De Angelis and L. Malavasi, Adv. Funct. Mater., 2021, 31, 2104428 CrossRef CAS .
  23. C. Barman, S. Lavadiya, S. K. Nayak, V. R. Soma and S. S. K. Raavi, ACS Appl. Nano Mater., 2025, 8, 2238–2248 CrossRef CAS .
  24. F. N. Mansoorie, P. Bhatt, A. Tewari, R. Kumar and M. Bag, ACS Appl. Mater. Interfaces, 2025, 17, 4218–4230 CrossRef CAS PubMed .
  25. Z. Ma, P. Lv, X. He, F. Wang, Y. Li, G. Xiao and B. Zou, Nano Lett., 2025, 25, 9345–9352 CrossRef CAS PubMed .
  26. O. A. Lozhkina, A. A. Murashkina, V. V. Shilovskikh, Y. V. Kapitonov, V. K. Ryabchuk, A. V. Emeline and T. Miyasaka, J. Phys. Chem. Lett., 2018, 9, 5408–5411 CrossRef CAS PubMed .
  27. L. Li and Z. Zhang, Chem. Eng. J., 2022, 434, 134811 CrossRef CAS .
  28. L. Protesescu, S. Yakunin, M. I. Bodnarchuk, F. Krieg, R. Caputo, C. H. Hendon, R. X. Yang, A. Walsh and M. V. Kovalenko, Nano Lett., 2015, 15, 3692–3696 CrossRef CAS PubMed .
  29. C. Lu, M. W. Wright, X. Ma, H. Li, D. S. Itanze, J. A. Carter, C. A. Hewitt, G. L. Donati, D. L. Carroll, P. M. Lundin and S. M. Geyer, Chem. Mater., 2019, 31, 62–67 CrossRef CAS .
  30. F. Ye, H. Zhang, W. Li, Y. Yan, J. Cai, R. S. Gurney, A. J. Pearson, D. Liu and T. Wang, Small Methods, 2019, 3, 1800489 CrossRef .
  31. F. Ye, H. Zhang, W. Li, Y. Yan, J. Cai, R. S. Gurney, A. J. Pearson, D. Liu and T. Wang, Ligand–Exchange of Low–Temperature Synthesized CsPbBr3 Perovskite toward High–Efficiency Light–Emitting Diodes, Small Methods, 2019, 3, 1800489 CrossRef .
  32. D. Lee, M. Kim, H.-Y. Woo, J. Chae, D. Lee, S. Jeon, S. J. Oh and T. Paik, RSC Adv., 2020, 10, 7126–7133 RSC .
  33. Z. Wang, C. Jiang, R. Huang, H. Peng and X. Tang, J. Phys. Chem. C, 2014, 118, 1155–1160 CrossRef CAS .
  34. D. Velasco-Arias, I. Zumeta-Dubé, D. Díaz, P. Santiago-Jacinto, V.-F. Ruiz-Ruiz, S.-E. Castillo-Blum and L. Rendón, J. Phys. Chem. C, 2012, 116, 14717–14727 CrossRef CAS .
  35. B. Wang, L. Liu, B. Liu, J. Li, B. Cao, Z. Zhao and Z. Liu, Phys. B, 2020, 599, 412488 CrossRef CAS .
  36. M. Laitz, A. E. K. Kaplan, J. Deschamps, U. Barotov, A. H. Proppe, I. García-Benito, A. Osherov, G. Grancini, D. W. deQuilettes, K. A. Nelson, M. G. Bawendi and V. Bulović, Nat. Commun., 2023, 14, 2426 CrossRef CAS PubMed .
  37. T. T. H. Do, A. Granados del Águila, D. Zhang, J. Xing, S. Liu, M. A. Prosnikov, W. Gao, K. Chang, P. C. M. Christianen and Q. Xiong, Nano Lett., 2020, 20, 5141–5148 CrossRef CAS .
  38. A. C. Jakowetz, M. L. Böhm, A. Sadhanala, S. Huettner, A. Rao and R. H. Friend, Nat. Mater., 2017, 16, 551–557 CrossRef CAS PubMed .
  39. G. Rey, G. Larramona, S. Bourdais, C. Choné, B. Delatouche, A. Jacob, G. Dennler and S. Siebentritt, Sol. Energy Mater. Sol. Cells, 2018, 179, 142–151 CrossRef CAS .
  40. S. N. Inamdar, P. P. Ingole and S. K. Haram, ChemPhysChem, 2008, 9, 2574–2579 CrossRef CAS PubMed .
  41. Q. Zhang, F. Diao, X. Xue, X. Sheng, D. Barba and Y. Wang, ACS Appl. Mater. Interfaces, 2021, 13, 44777–44785 CrossRef CAS PubMed .
  42. K. Marjit, G. Ghosh, R. K. Biswas, S. Ghosh, S. K. Pati and A. Patra, J. Phys. Chem. Lett., 2022, 13, 5431–5440 CrossRef CAS PubMed .
  43. B. Pereda-Ayo, U. De La Torre, M. J. Illán-Gómez, A. Bueno-López and J. R. González-Velasco, Appl. Catal., B, 2014, 147, 420–428 CrossRef CAS .
  44. V.-D. Quach, M. C. Spadaro, D. Dragoe, M. Botifoll, H. Vezin, C. Colbeau-Justin, F. Dumeignil, J. Arbiol, R. Wojcieszak and M. N. Ghazzal, J. Mater. Chem. A, 2024, 12, 19236–19246 RSC .
  45. Y. Dai and H. Tüysüz, Sol. RRL, 2021, 5, 2100265 CrossRef CAS .
  46. L.-Y. Pan, Y.-F. Ding, S.-F. Yin and M.-Q. Cai, New J. Chem., 2022, 46, 16922–16931 RSC .
  47. J. Li, X. Wang, H. Fang, X. Guo, R. Zhou, C. Wang, J. Li, M. N. Ghazzal and Z. Rui, J. Mater. Chem. A, 2023, 11, 25639–25649 RSC .
  48. J. Li, H. Fang, M. Wu, C. Ma, R. Lian, S. P. Jiang, M. N. Ghazzal and Z. Rui, Small, 2023, 19, 2300559 CrossRef CAS PubMed .
  49. R. Ciancio, R. E. Dunin-Borkowski, E. Snoeck, M. Kociak, R. Holmestad, J. Verbeeck, A. I. Kirkland, G. Kothleitner and J. Arbiol, Microsc. Microanal., 2022, 28, 2900–2902 CrossRef .

Footnote

These authors contributed equally.

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