Open Access Article
Andrea
Bartoletti
ab,
Angela
Gondolini
*a,
Nicola
Sangiorgi
a,
Matteo
Aramini
c,
Matteo
Ardit
d,
Marzio
Rancan
e,
Lidia
Armelao
bf,
Simon A.
Kondrat
*g and
Alessandra
Sanson
a
aInstitute of Science and Technology for Ceramics (ISTEC) of the National Research Council (CNR), Via Granarolo 64, I-48018 Faenza (RA), Italy. E-mail: angela.gondolini@istec.cnr.it
bDepartment of Chemical Sciences, Università degli Studi di Padova, Via Marzolo 1, 35131, Padova, Italy
cDiamond Light Source, Harwell Science and Innovation Campus, Chilton, Didcot, OX11 0DE, UK
dDepartment of Physics and Earth Sciences, University of Ferrara, Via Saragat 1, Ferrara 44122, Italy
eInstitute of Condensed Matter Chemistry and Technologies for Energy (ICMATE), National Research Council (CNR), c/o Department of Chemistry, University of Padova, Via F. Marzolo 1, 35131 Padova, Italy
fDepartment of Chemical Sciences and Materials Technologies (DSCTM), National Research Council (CNR), Italy
gDepartment of Chemistry, Loughborough University, Loughborough, Leicestershire, LE113TU, UK. E-mail: s.kondrat@lboro.ac.uk
First published on 21st December 2022
The perovskite CaCu3Ti4O12 is known for its ability to photocatalytically degrade model dye molecules using visible light. The influence of ball milling preformed CaCu3Ti4O12 on the catalysts structure and performance in the degradation of rhodamine B and the antihistamine cetirizine hydrochloride, which does not absorb light in the visible region, was investigated. The surface area of CaCu3Ti4O12 increased from 1 m2 g−1 to >80 m2 g−1 on milling with a retention of 96% CaCu3Ti4O12 phase purity, as determined by X-ray diffraction and extended X-ray absorption fine structure analysis. Multiple characterisation techniques showed an increase in structural defects on milling, including, for the first time, X-ray absorption near edge spectroscopy which showed changes in the local electronic structure from the perspective of Cu and Ti. Photocatalytic degradation was notably higher with the milled sample than that observed for the as-synthesized sample, even after normalisation for surface area, with a doubling of surface normalised rate constant from 4.91 × 10−4 to 9.11 × 10−4 L min−1 m2 for rhodamine B degradation and a tripling for cetirizine hydrochloride degradation from 2.64 × 10−4 to 7.92 × 10−4 L min −1 m−2. The improvement in catalytic performance can be correlated to the defects observed by X-ray absorption spectroscopy.
To find an alternative strategy to the inherently flawed doping of UV-active TiO2 with coloured transition metals, ordered mixed metal oxides with multiple photoactive cations have been employed for visible light driven photocatalysis.2,3 Rosseinsky and co-workers highlighted the potential of this strategy with a CaCu3Ti4O12 double-perovskite catalyst for visible light photooxidation of model pollutants.4 CaCu3Ti4O12 is unusual, in that in addition to Ca, it has square planer Cu2+ on the A site, accommodated by the significant tilting of the vertex sharing TiO6 octahedra. The material is well known for its dielectric properties,5 photoluminescence6 and antiferromagnetism.7 Its potential as a visible light photocatalyst being due to photoexcitation from occupied CuO states at the top of the valance band to a Ti based conduction band, reducing the band gap relative to TiO2 alone. However, the photoexcitation process is partially hindered by localised Cu(3d)-O(2p) σ-antibonding states within the band gap, that limit the mobility of Cu+ carriers thus facilitating carrier recombination.4 Further, the conventional synthesis of perovskites (e.g., solid-state reaction synthesis) commonly produces compounds with low surface area (ca. <5 m2 g−1), and therefore a low number of surface-active sites.
Since this seminal work, several reports have focused on synthesising the CaCu3Ti4O12 photocatalyst and applying it to a host of other organic substrates for photodegradation. Attempts to overcome recombination and surface area issues have been made by producing hybrid catalysts, including incorporation of Pt nanoparticles,4 polymers,8,9 carbon nanotubes,10 carbon nitrides11 and additional metal oxides (zeolites, SiO2 and TiO2).12,13 Other work has utilised alternative sol–gel and molten salt routes to improve surface area, control morphology and induce defects in the structure.14–17
The presence of point or planar defects such as oxygen nonstoichiometry, metal site vacancies or antisite disorder, are often reported and hypothesised to influence CaCu3Ti4O12 properties. For example, the unexpectedly high dielectric constant of CaCu3Ti4O12, relative to that predicted by DFT, has been hypothesized as being due to planar defects or Ca–Cu antisite defects.18,19 While, photoluminescence effects are ascribed to [TiO5VO]–[TiO6] clusters from oxygen vacancies.6 Photocatalytic properties for organic substrate decomposition have been stated to be improved by bulk and surface defects, as determined by photoluminescence measurements, EPR and XPS. Hailili et al.14–16 stated that the photocatalytic performance of CaCu3Ti4O12 could be correlated with EPR line broadening, which has been attributed to oxygen vacancies and other defects, such as the formation of Cu+[CuO4]′ and Ti3+[TiO5VO]′ species. Complimentary XPS analysis further showing additional features in the Cu 2p and O 1s that were attributed to the presence of Cu1+ on the catalyst surface and oxygen vacancies. These defects are theorised to alter carrier mobility and act as electron scavenger sites to generate superoxide radicals to the benefit of catalytic activity.
While molten salt synthesis routes have produced defined morphologies with structural defects, surface areas remain below 30 m2 g−1.16 Variation of Cu–Ca ratios in conventional oxide synthetic routes has also been used to create defective CaCu3Ti4O12 but at the expense of phase purity (CuO and CaTiO3 being formed) and poor surface area. Mechanical milling of CaCu3Ti4O12 represents an appealing strategy to increase surface area substantially, retain phase purity (if the process is sufficiently controlled), and also induce structural defects. The latter point being highlighted in a series of papers by Modi and co-workers,20,21 where on milling, positron annihilation spectroscopy and photoluminescence measurements confirmed an increase in structural defects, UV-vis spectroscopy that mechano-chromism can be induced, and lastly that antiferromagnetic properties are disrupted.
Therefore, the impact of the simple process of milling on the photocatalytic properties of CaCu3Ti4O12 requires further investigation. The prospect of a simple and affordable process to high surface area mixed metal oxide photocatalysts being appealing. Thus, this study describes the mechanochemical processing of CaCu3Ti4O12, which provides the highest reported surface area of this material to date and exceptional mass and surface area normalised photodegradation rate of rhodamine B and cetirizine hydrochloride (Ctz). These substrates have been chosen, as they have been frequently employed for studying the photocatalytic properties of metal oxides, including titania and also CaCu3Ti4O12, allowing for effective comparison with other synthesis methodologies within the literature. The issue of rhodamine B dye sensitization in its photodegradation masking or complicating structure–function relationships, which is frequently not discussed, has been mitigated through the use of the second Ctz substrate that does not absorb light in the viable region. Structural defects, which are important to catalytic performance, have been studied by X-ray absorption spectroscopy and supported by theoretical simulations.
:
3) were dissolved in deionized water and absolute ethanol (3
:
1 vol%) to produce a 2.0 M nitrate solution. Meanwhile, stoichiometric amount of Ti(OC4H9)4 was dissolved in ethanol to obtain a 0.5 M solution. Once the complete precursor dissolution occurs the as-produced solutions were mixed until the formation of a sky-blue gel. The as-obtained gel was heated up to 60 °C for 4 h in a water bath and then at 100 °C for 24 h to obtain the dried precursors. The produced powder was homogenised for 30 minutes in agate mortar, calcined at 1100 °C for 2 h in air and sieved at 64 μm to obtain the final CaCu3Ti4O12 powder (CCTO). Part of the obtained powder was milled in a high energy planetary mill (Fritsch, Pulverisette 6) using zirconia beads as grinding media and absolute ethanol as solvent. The milling speed was kept at 400 rpm, using a powder to beads ratio of 1
:
10 until a cumulative kinetic energy of 4.6 × 104 kJ g−1 was conferred to the system (the description of the calculation method is reported in ESI,† Section S1, Table S1). The milled powder was dried at room temperature and sieved at 64 μm to obtain the so-labelled CCTO-GM sample.
Specific surface area of powders was obtained from Brunauer, Emmett, and Teller analysis of nitrogen adsorption isotherms in vacuum with a Sorpty 1750 (Carlo Erba Strumentazione). Before analysis, all the samples were degassed at 100 °C for 1 h. Morphology of the as-prepared and milled samples was investigated using a ∑IGMA SEM-FEG(Zeiss).
XPS analyses were performed with a Perkin-Elmer Φ 5600-ci spectrometer using Mg Kα radiation. The sample analysis area was 800 μm in diameter. The standard deviation for the BEs values was ± 0.2 eV (further details are reported in ESI,† Section S2). It cannot be excluded that Cu+ in CCTO can derive from X-ray irradiation or vacuum effects during XPS analysis.23,24 For these reasons, particular attention has been posed to assure the same experimental conditions for the two samples (i.e., time in vacuum, X-ray source power, and X-ray exposure time).
Photoluminescence (PL) analysis where performed using a Horiba JobinYvon Fluorolog-3 spectrofluorimeter in front-face acquisition geometry exciting samples at 380 nm.
X-ray absorption spectroscopy was performed on the B18 beamline at Diamond Lightsource. Measurements were performed using the Si (111) monochromator in both transmission and fluorescence mode. Samples were diluted in cellulose and pressed into pellets to achieve the optimal edge step. All data was calibrated using appropriate Cu and Ti foils. X-ray absorption spectroscopy and EXAFS processing was performed using Athena and Artemis software. Amplitude reductor factors were determined from known CuO and TiO2 reference compounds. Aside from 1st shell M–O paths, all coordination numbers were fixed at those expected from the crystallographic data and mean squared displacement parameters floated during fits.
Infrared spectroscopy experiments were carried out on investigated powders afters photocatalytic measurements using a Nicolet™ iS™ 5 FTIR spectrometer (Thermo Fischer instrument) equipped with an internal reflection element (ATR crystal) i.e., a diamond crystal. Collected spectra, obtained by means of 16 scans in the range of 400–4000 cm−1, have a resolution of 4 cm−1.
In the photolysis process, the electronic band structure of each sample has to be considered. The position of the conduction (ECB) and the valence (EVB) bands of a semiconductor can be calculated from eqn (1) and (2) respectively:14,15
![]() | (1) |
| EVB = Eg + ECB | (2) |
Transmission electron microscopy (TEM) was performed on a JEOL JEM 2100 LaB6 operating at 200 kV. Samples were dry dispersed onto holey carbon film 300 mesh copper grids.
![]() | (3) |
![]() | (4) |
An extended (10 × 10 × 10) k-point grid was instead used for the calculation of the X-ray absorption spectra. Self-consistent calculations were performed to a energy threshold of 1 × 10−7 eV. Due to the isolated nature of the core–hole excitation, a 2 × 2 × 2 supercell bearing P1 symmetry was used for the calculation of the excited state, using the same convergence parameters as found for the ground-state calculation, after evaluating that the convergence criteria were still respected, and that the choice of larger values was not influencing both the energy and the simulated spectra.
An energy shift of 8979 eV was applied to the simulated spectra, in order to overlap with the experimental data, and computational results were normalised through trapezoidal integration of simulated spectrum. Transition broadening as a consequence of instrumental resolution (Gaussian) and core-lifetime effects (Lorentzian) was applied as 0.2 and 1.25 eV FWHM respectively.
:
3
:
4. As expected but often not reported, high energy milling resulted in the sample contaminated with 3 ppm of Zr from the mill media, which was considered sufficiently low in concentration to have minimal influence on catalytic properties.
| CaCu3Ti4O12 | prv | ten | ||||||||
|---|---|---|---|---|---|---|---|---|---|---|
| R exp | R wp | G.O.F | wt% | a (Å) | V (Å3) | D XRD (nm) | Strain × 10−4 | wt% | wt% | |
| CCTO | 0.022 | 0.034 | 0.015 | 95.6(1) | 7.393(2) | 404.02(3) | 263(4) | 0.5(1) | 1.9(1) | 2.5(1) |
| CCTO-GM | 0.022 | 0.033 | 0.015 | 95.7(1) | 7.395(1) | 404.32(2) | 22(2) | 1.8(1) | 3.2(1) | 1.1(1) |
A notable effect of milling was a reduction in CaCu3Ti4O12 crystallite sized determined from XRD, from 260 to 20 nm for the CCTO and CCTO-GM, with the associated reduction in particle size being confirmed by SEM and TEM analysis (Fig. 2 and S1†). TEM analysis shows the presence of large crystallites in the CCTO sample (Fig. 2a and b) with the CCTO-GM displaying agglomerations of smaller crystallites without a distinct morphology after milling (Fig. 2c and d). The inset selected area electron diffraction highlights the transformation from large single crystals to a polycrystalline material. Such a reduction in particle size on milling is reflected in a significant increase in surface area of the powder from 1.4(1) m2 g−1 to 80.1(4) m2 g−1. The reported surface area of CaCu3Ti4O12 after milling is, to the authors knowledge, the highest reported in the literature and significantly greater than those reported using sol–gel and molten salts synthetic routes (<30 m2 g−1).4,14–16 It is anticipated that such an increase in surface area will significantly enhance photocatalytic performance through increasing the number of adsorption sites (vide infra).
![]() | ||
| Fig. 2 Transmission electron micrographs of CCTO (a and b) and CCTO-GM (c and d). Inset: selected area electron diffraction patterns of the areas shown. | ||
While the CaCu3Ti4O12 phase is retained on milling, the observed lattice parameter from XRPD increased from 7.393 (CCTO) to 7.395 Å (CCTO-GM) and it was coupled with an increase in crystallite microstrain. The latter situation has previously been observed by Raval et al.,20 and was attributed to milling induced defects and lattice disorder at grain boundaries. As stated above, the presence of defects has previously been reported to influence the photocatalytic performance of CaCu3Ti4O12 when prepared using molten salts. Therefore, it is important to further understand the nature of defects induced by milling and how they influence the intrinsic catalytic performance of CaCu3Ti4O12 in addition to the potential benefit from an increase in surface area.
| CCTO | ||||
|---|---|---|---|---|
| Transition | Energy (eV) | Max. band wavelength (nm) | E CB (V vs. NHE) | E VB (V vs. NHE) |
| α | 1.8 | 681 | 0.068 | 2.15 |
| β | 2.1 | 596 | ||
| γ | 4.2 | 293 | ||
| CCTO-GM | ||||
|---|---|---|---|---|
| Transition | Energy (eV) | Max band wavelength (nm) | ECB (V vs. NHE) | EVB (V vs. NHE) |
| α | 1.8 | 692 | −0.044 | 2.26 |
| β | 2.3 | 538 | ||
| γ | 4.2 | 298 | ||
Milling of CaCu3Ti4O12 (CCTO-GM) resulted in an appreciable blue shift and increase in intensity of the band related to transition β. Further, the intensity of feature γ significantly increased on milling. Band gaps from direct transitions of 2.1(1) eV and 2.3 (1) eV are derived from Tauc plots (Fig. 3b) for CCTO and CCTO-GM samples, respectively. It is well-known that the optical and electronic properties of materials heavily depend on the synthesis method and subsequent powder treatments, as well as on the physical state of the samples under examination (powder, pellets, films). Comparable blue shift and relative intensity changes have been observed by P. Y. Raval et al. on the milling of CaCu3Ti4O12, which were attributed to a lowering in the point symmetry of [CuO4] units from D2h (square planar coordination) to D2 (highly distorted tetrahedral coordination), combined with a reduction in particle size.21 Yet, no other evidence is provided to support such a change in Cu local coordination (note such as change is not supported from Cu XANES vide infra). Hailili et al. also saw significant changes in spectra for CaCu3Ti4O12 prepared using different molten salt synthesis techniques.14,15 The intensity of β features were associated with metal site defects (i.e., Ti3+ and Cu–O vacancies), supported by EPR and XPS analysis. However, the observed changes in UV-vis spectra cannot easily be attributed solely to defects and can be influenced by a number of other parameters, such as particle size and impurities.
:
Ti4+ ratio of 1
:
10. The fitting of the Cu2p3/2 region, along with its shake-up feature, has been performed using three peaks to model the Cu2+ component, with a single Cu+ contribution in the main peak as previously reported.32,33 As seen in Fig. 4, both samples have Cu2+ and Cu+ contributions, with the latter having comparable BEs of 932.1 eV to that reported previously for other CCTO samples.32 However, CCTO-GM surface had a higher concentration of Cu+ than the unmilled CCTO, with a Cu2+
:
Cu+ ratio of 3
:
1 and 2
:
1 in CCTO and CCTO-GM, respectively. Increased signals fitted as Cu+ and Ti3+ from XPS have, as discussed above, previously been used to validate the hypothesis of increasing oxygen defects within CaCu3Ti4O12. Our observations concur that the increase of structural defects, identifiable from changes in the optical properties, paired with the increase of crystallite microstrain and the lattice parameter, can be correlated with the increase of Cu+ and Ti3+ as detected from XPS analysis. Yet, several points of caution require highlighting, namely: (i) XPS is a surface technique and an over-extrapolation to bulk properties must be avoided. For instance, structural features at the perovskite surface are known to significantly differ to that of the bulk.34 (ii) Analysis of Cu oxidation state from Cu2p region is challenging due to small shifts in binding energy coupled with the broadness of the Cu2+ 2p3/2 peak.
It is also worth noting that the CCTO-GM XPS spectrum revealed the presence of Zr with a Zr/Ti ratio of 0.24. Notably, the latter value is significantly higher than that of the bulk, where a Zr/Ti ratio of 0.03(1) was found, meaning that Zr species are not well dispersed within the bulk of CCTO-GM, but they preferentially concentrate on the CCTO surface as small particles. The Zr 3d5/2 binding energy was observed at 182.3 eV and suggested that the Zr was present in an oxide environment such as ZrO2,35i.e. fragments of milling media.
| Sample | Path | Coordination number(i) | Symmetric expansion | Path length (Å) | 2σ2 (Å2) | ΔE (eV) | R-Factor |
|---|---|---|---|---|---|---|---|
| Fitting parameters: 3 < k < 12 and 1.1 < R < 3.7. (i) S02 values (fitted from CuO) fixed at 0.85. (ii) Cu–O(1) increase to 3.7(1) if 2σ2 fixed at 0.0048. (iii) 2σ2 defined as 2σ2 for Cu–O(2) × 1.17. (iv) 2σ2 defined as 2σ2 for Cu–O(2) × 3.6. Multiples defined by factor of increased path length vs. path length of calculated 2σ2. | |||||||
| CCTO | Cu–O(1) | 3.5(2)(ii) | n/a | 1.953(3) | 0.0042(1) | −0.2(6) | 0.0034 |
| Cu–O(2) | 4 | 0.0025(20) | 2.788 | 0.0184(36) | |||
| Cu–Ti | 8 | 3.208 | 0.0080(4) | ||||
| Cu–O(3) | 4 | 3.271 | 0.0216(iii) | ||||
| Cu–O–Ti–Cu (MS) | 16 | 3.573 | 0.0204(iv) | ||||
| Cu–Cu | 4 | 3.705 | 0.0096(14) | ||||
| Cu–Ca | 2 | 3.705 | 0.0111(47) | ||||
| CCTO-GM | Cu–O(1) | 4.3(2) | n/a | 1.955(2) | 0.0048(6) | 0.3(5) | 0.0026 |
| Cu–O(2) | 4 | 0.0057(15) | 2.797 | 0.0179(26) | |||
| Cu–Ti | 8 | 3.218 | 0.0093(4) | ||||
| Cu–O(3) | 4 | 3.281 | 0.0209(iii) | ||||
| Cu–O–Ti–Cu (MS) | 16 | 3.584 | 0.0149(iv) | ||||
| Cu–Cu | 4 | 3.716 | 0.0107(13) | ||||
| Cu–Ca | 2 | 3.716 | 0.0166(64) | ||||
Considering further shells, differences can be observed between samples in the magnitude of the Fourier Transform of the k3 weighted EXAFS, with a notable decrease in the feature just below 3 Å (not phase corrected). This feature is dominated by the Cu–Ti path with contributions from a Cu–O–Ti–Cu multiple scattering path and the Cu–Cu/Cu–Ca paths. The fitting was performed with fixed CN values with the 2σ2 values being used as an indicator of structural disorder (thermal disorder being constant). A clear increase in 2σ2 can be determined for the Cu–Ti path on milling, demonstrating a reduction in local structural order/CN. Reduced coordination is often seen for metal–metal paths in nanostructured oxides, due to the strong contribution of undercoordinated surface sites, however this is normally observed for particles <5 nm. Therefore, the observed particles are considered too large from an EXAFS perspective to see significant particle size affects. Numerous studies have suggested that oxygen vacancies and the formation of distorted [TiO5VOz] square pyramidal geometries can be accommodated in CaCu3Ti4O12. Whangbo and Subramanian hypothesized that sterically crowded CuO6 units, formed by twinning parallel to the (100) plane, could relax to form CuO4 units through oxygen vacancy formation on the connected distorted TiO6.18 The electrons left to the plane defect then facilitate Cu+ and Ti3+ formation. The increased structural disorder in the Cu–Ti path on milling may reflect such planer defects being formed. Increases in 2σ2 were also observed on milling for both Cu–Cu and Cu–Ca paths. Zheng et al. demonstrated that Cu–Ca antisite defects are present in CaCu3Ti4O12, through a combination of techniques including Cu K-edge EXAFS.19 Relatively high 2σ2 values for these paths (0.0096 and 0.0111 Å2) could be attributed as evidence for antisite defects. However, the error in fitted 2σ2 values makes change in the contribution of these defects on milling difficult to verified.
Ti K-edge EXAFS (Fig. 6) proved challenging to fit, with high errors in the fit preventing clear distinction between CCTO and CCTO-GM (Fig. S2, S3 and Table S3†). A clear difference in phasing and the magnitude of the FT between CCTO and TiO2 showed that there are no significant amounts of TiO2 in either CaCu3Ti4O12 sample. Further, both samples could be fitted to a CaCu3Ti4O12 model, although Ti–O distances were slightly shorter than predicted (1.94(1) Å vs. an expected 1.96 Å), potentially demonstrating reduced Ti–O coordination as reported in simulations of TiO2 nanostructures. Between CCTO and CCTO-GM differences are observed in the intensity of features. In particular, the feature at 2.8 Å in the magnitude of the FT, which has contributions from Ti–Cu/Ca paths (in addition to Ti–O multiple scattering, Ti–O (2nd shell) and Ti–Ti paths), is dampened in CCTO-GM. The 2σ2 values for the Ti–Cu/Ca paths were higher for CCTO-GM at 0.009(2)Å2vs. 0.006(2) Å2 seen for CCTO. Broadly, and without over interpreting the K-edge EXAFS data, this correlates with the observations from the Cu K-edge data, in that the A–B (Cu–Ti) site path is more disordered after milling.
![]() | ||
| Fig. 6 Ti K-edge k2χ data EXAFS (left) and the associated magnitude of the Fourier transform (right) of CCTO, CCTO-GM and rutile TiO2. (Black) CCTO; (blue) CCTO-GM; (dashed red) rutile TiO2. | ||
![]() | ||
| Fig. 7 Cu K-edge XANES of CCTO and CuO. (Black) CCTO, (blue) CCTO-GM and (red dashed) CuO. (A) 1s → 3d, (B) 1s → 4pπ, (C) 1s → 4pσ, (*) shakedown feature. | ||
Solomon and co-workers attributed the relative intensity of peak B*/B with the degree of bond covalency,38 while Heald et al. noted lower than expected B feature in La1.85Sr0.15CuO4, and attributed this to the presence of holes in the O 2p band.39 Hence, simplistically it can state that milling has created vacancies or reduced covalency of the Cu–O bond, which limits the probability of ligand to metal transfer and thus the screening of this transition. To shed further light on this, ab initio simulations were performed to provide XANES spectral positions, intensities and the wave-function associated with the orbital response of any given band. Fig. 8a shows that the Cu k edge XANES could be simulated with good agreement to experiment. From simulation, feature B* at 8.988 keV has an excited state with O hybridised with Ti d states, while B is localised on the solely Cu–O hybridised orbitals. Lastly, peak C at 8998.5 eV, responsible for the strongest feature in the spectra, is the result of the collective contribution of p-symmetry states of two neighbouring Cu sites, linked together by O p states. With this additional information, the reduction of the B* feature on milling suggests a reduction in the interaction between Cu states and Ti(d)–O(p) hybridised states, potentially through disruption of Cu–Ti centres ordering as seen from EXAFS.
Ti K-edge XANES, shown in Fig. 9, provides analogues evidence, from the perspective of Ti of a decrease in local ordering. Detailed comparison between theory and experiment allows for a correlation of Ti coordination number with band position and intensity (increasing intensity and lower energy position = lower coordination).6,40 The complex pre-edge features, seen in all samples, correlate with those seen by Oliveira et al.6 and were assigned as; (A) 1s → t2g of [TiO6] octahedra quadrupole excitations, (B) 1s → eg of [TiO6] octahedra and (C) dipole excitation 1s → eg of neighbouring [TiO6] octahedra. Notably, on milling, the position of feature B shifts from 4971.2 eV to 4970.8 eV, while the normalised peak area slightly increases. Using the correlation determined by Farges and co-workers it can be estimated that the average Ti coordination number decreased on milling.40 Oliveira et al. attributed this to the formation of oxygen vacancies as [TiO5VOz], where Z denotes different charged vacancies, and therefore the coordination number can be expressed as a [TiO6]
:
[TiO5VOz] ratio. The ratio for CCTO can be determined as being 4.2
:
1 [TiO6]
:
[TiO5VOz], which then decreases on milling (CCTO-GM) to 1.2
:
1. In addition to the change in feature B, the intensity of features C decrease on milling, which suggests that the interaction between adjacent vertex shared [TiO6] has changed. The change can be hypothesised as being a subtle alteration in the tilting angle of the octahedra or due to vacancies, leading to a reduced excitation 1s → eg of neighbouring [TiO6] octahedra. These findings are supported by changes in the photoluminescence spectra (Fig. S4†) of the two samples, with a splitting of the predominant peak at 400 nm and an decrease in intensity on milling. These changes are interpreted as an increase in the concentration of oxygen vacancies and a change in their nature within complex [TiO6–TiO5VOz] clusters on milling, although changes in scattering on milling could also contribute to intensity changes. Lastly, Farges and co-workers noted that the intensity of main features (M1 and M2) broaden and reduce in intensity with medium range disorder (i.e. glassy vs. crystalline states).40 Interestingly, the normalised intensity of these features didn't decrease on CCTO milling but marginally increased, which suggests a retention, or moderate increase, of medium range order.
In summary, detailed characterisation shows that CaCu3Ti4O12 retains medium and long-range order on milling with an increase in crystallite strain coupled with an increase in surface area due to a reduction in crystallite/particle size. EXAFS supports the retention of the CaCu3Ti4O12 phase but with a decrease in ordering between A and B centres. Cu K-edge XANES showed that milling resulted in a change in the electronic structure, with a reduced probability of transitions to an excited state with significant Ti d state character. An increase in [TiO5VOz] concentration, coupled with a change in short range [TiO6] octahedral interactions is observed. Currently this is hypothesised as being due to an increase in plane defects and proportion of surface structures on milling. These defects are responsible for the observed change in optical properties with little evidence for a significant change in Cu coordination environment.
Fig. 10 shows the conversion of rhodamine B and the related first order kinetic decay plots for both catalysts under their respective ideal conditions. As a control, rhodamine solution was illuminated using visible light irradiation in absence of photocatalysts with no change in concentration being observed, indicating that degradation takes place by the photocatalytic action rather than by self-photosensitization.4,8,22 CCTO and CCTO-GM samples show notable adsorption capacity towards rhodamine B before exposure to sunlight, as previously reported by K. V. Ivanov et al.41 After this initial absorption no further reactivity was observed, with blank experiments in dark conditions show no remarkable rhodamine B conversion between 60 and 140 minutes. Under solar irradiation the rate constants for the two catalysts were distinctly different at 2.75 × 10−3 min−1 and 1.02 × 10−2 min−1 for CCTO and CCTO-GM respectively. As anticipated the significant increase in surface area on milling was beneficial to catalytic performance. Further, when normalised to surface area and catalyst concentration the rate of the milled CCTO-GM was almost twice that of the original perovskite at 9.11 × 10−4 L min −1 m−2 and 4.91 × 10−4 respectively. Therefore, milling intrinsically increased catalytic activity, which can be ascribed to the influence of structural defects, as identified by changes to XRD determined unit cell parameters, Cu–Ti path disorder from Cu K-edge EXAFS and changes in the Cu and Ti K-edge pre edge XANES. These are considered to be line or point defects but may also be associated with a higher proportion of high index surface terminations on milling. A further consideration is that an intrinsic particle size effect is responsible for improved reactivity, with CCTO-GM having significantly smaller particle size than CCTO. i.e. that the formation of nanoparticles has resulted in a change in band structure and hence photocatalytic performance. However, it should be considered that the average crystallite size of CCTO-GM from XRD is 20 nm, which is considerably larger than that expected for such changes in band structure.
To better understand the dye adsorption behaviour on the catalyst's surface, Z potential analysis were carried out at pH 7 (pH of the RhB solutions before degradation). At this pH the dye can exist in a zwitterionic form [29] and can interact with the powders through the negatively (COO−) or positively (NR4+) charged functionalities. CCTO and CCTO-GM samples show both negative surface charge of −82(3) mV and −31(9) mV, respectively and, therefore, can interact with the cationic part of the dye molecules. Thus, the adsorption difference found between the two samples are probably linked to the surface area of the powders combined with the surface chemistry, that are both greatly influenced by the milling treatment.
N+(C2H6) of the rhodamine B structure (see ESI† Section S3, Fig. S8, Table S4). No changes were seen in this adsorbed species after photocatalytic reaction.
![]() | ||
| Fig. 11 IR spectra of CCTO (a) and CCTO-GM (b) samples as-prepared, after an hour of stirring with rhodamine B in dark and after photodegradation. | ||
CCTO-GM in contrast to CCTO had clear features in the as-prepared sample, which could be assigned to surface bound carbonate (1396 cm−1 and 1556 cm−1) and adsorbed water (1650 cm−1 and 3400 cm−1). The presence of carbonate in CCTO-GM was also confirmed from the XPS C1s signal (Fig. S7†). The significant splitting of the ν(CO) mode indicates a significant loss of symmetry from the free CO32− anion (D3h) and that the carbonate is strongly bound to surface metal sites. After exposure to the substrate in the dark, CCTO-GM had two absorption bands at 1149 and 1225 cm−1 ascribable to C–O–H stretching of carboxylic group,8 and C–N stretching vibration of the ammino group presents in RhB. These two peaks were shifted to higher wavenumbers compared to the as prepared sample indicating the formation of weaker interactions through the milled sample and RhB molecules. This can be related to the different surface chemical environment (presence of carbonates and zirconium, higher Cu+/Cu2+ and Ti3+/Ti4+ ratios and a more positive surface charge) of CCTO-GM compared to the as prepared sample. Moreover, the bending of C–H bond in the group
N+(C2H6) (visible at 1400 cm−1 in the CCTO spectra) cannot be appreciated in CCTO-GM probably due to the overlap with the carbonate's modes centred at 1396 cm−1. FTIR results suggest that rhodamine B molecules are firstly adsorbed onto calcium copper titanate's surface (a schematic rapresentation and description is presented in ESI† Section 5 Fig. S10) and subsequently degraded under sun light irradiation through the destruction of the chromophore group.8
Moreover, the addition of pBQ seems to reduce the photodegradation rate, suggesting that photoproduced O2˙− radicals are involved in dye degradation mechanism. The presence of DMSO does not have a great influence on the photodegradation reaction, indicating that photoproduced electron are not the main reactive species during dye degradation. Using the band edge positions calculated with eqn (1) and (2) (Table 2), a schematic illustration of the possible photodegradation mechanism is reported in Fig. 13. All standard potentials are taken at pH 7.
![]() | ||
| Fig. 13 Schematic representation of solar light driven degradation mechanisms of rhodamine B by a) CCTO, b) CCTO-GM. | ||
Both samples show the edge of the valence band more negative than E°(H2O/OH˙) (2.730 V vs. NHE)45 but more positive than E°(OH−/OH˙) (1.902 V vs. NHE).45 Thus, the formation of hydroxyl radicals is thermodynamically favourable from the oxidation of adsorbed OH− groups, but not directly from water molecules. OH˙ and h+ photogenerated are the main active species during RhB photodegradation in CCTO-based compounds12,46 as generally reported for photodegradation of dyes and organic pollutants.47 The position of the conduction band is more positive than E°(O2(aq)/O2˙−) (−0.160 V vs. NHE)45,48 in both CCTO and CCTO-GM, therefore it would be difficult for these radicals to participate in the degradation of pollutant molecules. However, the presence of Ti3+ species in CCTO-band gap,4,14,15 as a consequence of the reaction of the electron transferred between Ti4+/Ti3+ (E° = −0.498 V vs. NHE) and the oxygen molecules adsorbed on the material surface, to give O2˙− species following the scheme:
| Ti3+ + O2 → Ti4+ + O2˙− |
Importantly, but often negated, is the possible importance of dye sensitization process in photocatalytic dye degradation. The energy of the highest occupied molecular orbital and lowest unoccupied molecular orbital (HOMO and LUMO) in rhodamine B are 1.1 and −1.0 V vs. NHE.51 Thus, a photo-excited electron can be transferred from the LUMO of RhB* to the more positive conduction band of CCTO-type catalyst, lowering the photoproduced charge recombination rate and enhancing dye degradation.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2cy01299e |
| This journal is © The Royal Society of Chemistry 2023 |