Open Access Article
Anika Rahman Riya†
a,
Abrar Daiyan†a,
Mehedi Hasan Prince†ab,
Troyee Mitra Aishi†ac,
Tinni Dey Promea,
Md. Abdullah Zubaird,
Md. Fakhrul Islamd and
Takian Fakhrul
*a
aDepartment of Materials and Metallurgical Engineering, Bangladesh University of Engineering and Technology (BUET), Dhaka, 1000, Bangladesh. E-mail: takianf@mme.buet.ac.bd
bDepartment of Materials Science & Engineering, Rensselaer Polytechnic Institute (RPI), NY 12180, USA
cDepartment of Chemical Engineering, Northeastern University, MA 02115, USA
dDepartment of Nanomaterials and Ceramic Engineering, Bangladesh University of Engineering and Technology (BUET), Dhaka, 1000, Bangladesh
First published on 15th October 2025
The growing environmental threat posed by synthetic dye pollution has accelerated the search for effective semiconductor-based photocatalysts for scalable and sustainable wastewater treatment. This study reports the greatly enhanced visible-light photocatalytic activity of Bi0.97Ca0.03Fe1−xCrxO3 (x = 0.00, 0.01, 0.03, 0.05) nanoparticles synthesized via a modified sol–gel method. Structural analysis confirmed the formation of a distorted rhombohedral perovskite phase, with doping leading to reduced lattice parameters and crystallite size. The bandgap narrowed from 2.13 eV in pure BiFeO3 (BFO) to 1.80 eV in the co-doped sample. Elemental analysis confirmed improved stoichiometric balance and reduced bismuth volatilization in the co-doped samples. Under xenon lamp illumination, methylene blue degradation reached 93% for Bi0.97Ca0.03Fe0.95Cr0.05O3 at neutral pH, approaching the upper limits of reported efficiencies for BFO-based systems. The reaction followed pseudo-first-order kinetics, with the rate constant rising from 0.01358 to 0.03038 min−1. Moreover, the formation of an antiferromagnetic–ferromagnetic core–shell structure in the co-doped samples is proposed to notably improve dye degradation by promoting surface charge separation and suppressing recombination. This work highlights the potential of Ca and Cr co-substituted BFO nanoparticles as high-performance, visible-light-driven photocatalysts for dye remediation under environmentally relevant pH conditions.
In this context, bismuth ferrite has emerged as a promising candidate due to its relatively narrow Eg (∼2.2 eV), which enables effective absorption of visible light.16,17 Its magnetic properties, arising from the breakdown of antiferromagnetic ordering, along with its chemical stability, facilitate easy recovery and reuse after degradation.18,19 Moreover, due to its multiferroic nature, the internal electric field promotes charge separation and reduces recombination, making it ideal for photocatalytic applications.20–22
Despite its potential, the photodegradation efficiency of BFO remains inconsistent across studies, largely influenced by synthesis methods, surface chemistry, and pH conditions. Extensive studies have explored the effect of solution pH on BFO's photocatalytic activity, with many reporting enhanced degradation efficiencies under both acidic and basic conditions, deviating from the dye's natural pH.23–27 The influence of pH is particularly relevant for dyes like methylene blue (MB), a persistent and carcinogenic cationic pollutant that degrades more efficiently in non-neutral environments.23,25 In parallel, other studies have shown that compositional tuning of the photocatalyst can improve degradation performance without the need for pH adjustment.22,28–31 Incorporating suitable elements into the BFO lattice introduces charge-trapping centers that enhance carrier mobility and mitigate recombination losses.32 For instance, Ca2+ doping induces structural distortions, generates oxygen vacancies, and modulates the band structure. These changes improve charge separation and photocatalytic activity.33,34 Similarly, Cr doping introduces trap states and increases radical generation under visible light, further boosting degradation efficiency.35 Co-substitution with both Ca and Cr enhances the magnetic properties of BFO while inducing the formation of an antiferromagnetic–ferromagnetic (AFM–FM) core–shell structure. This configuration facilitates efficient charge transport and allows for magnetic recovery of the catalyst after treatment.26,37,38 Investigating this combined doping approach may provide valuable insights into optimizing BFO-based systems for visible-light-driven environmental remediation.
Building on this foundation, this study optimizes the photocatalytic degradation of methylene blue by tuning BFO's bandgap through targeted doping, while maintaining the solution at pH 7 to isolate compositional effects. Pure, Ca-doped, and Ca–Cr co-substituted BFO nanoparticles (NPs) with compositions BiFeO3 and Bi0.97Ca0.03Fe1−xCrxO3 [(BCFO; x = 0), (BCFOCr1; x = 0.01), (BCFOCr3; x = 0.03), (BCFOCr5; x = 0.05)] were synthesized via the sol–gel method. Characterization included X-ray diffraction spectroscopy (XRD), scanning electron microscopy (SEM), transmission electron microscopy (TEM), UV-visible spectroscopy (UV-vis), and photocatalytic degradation tests using MB dye, were employed to investigate the structural, morphological, optical, and photocatalytic properties of the synthesized nanoparticles.
:
1 molar ratio, were dissolved in deionized water and stirred at 90–95 °C for 30 minutes. Subsequently, 10 mL of ethylene glycol was added to the solution, which was then stirred at 75–85 °C for 4 hours to induce gel formation. The resulting gel was dried at 110 °C for 24 hours, ground into a fine powder, and annealed at 550 °C for 2 hours with a heating rate of 5 °C min−1 to obtain crystalline nanoparticles.
The photocatalytic response of the nanoparticles was evaluated using an in-house-designed photochemical reactor rig. The bandgap of the samples was investigated using a UV-Vis spectrophotometer (LAMBDA 1050, PerkinElmer). Photodegradation of methylene blue solution was carried out with the nanoparticles as the photocatalyst, under illumination from a concentric 500 watt xenon lamp emitting photons in the 250–1800 nm range. The temperatures of the source chamber and the solution were regulated using forced air circulation and a closed-loop water cooling system, respectively, maintaining the solution within 23–26 °C. A homogeneous nanoparticle suspension was ensured by magnetic stirring at 1000 rpm, and the pH of the solution was maintained at approximately 7. A total of 100 mg of the nanoparticle catalyst was added to 350 mL of a 3.2 ppm MB dye solution. The mixture was stirred in the dark for 30 minutes to establish adsorption–desorption equilibrium, followed by 90 minutes of illumination. Aliquots of 15 mL were collected at 10 minute intervals, and the absorption spectra of the centrifuged samples were recorded using a UV-Vis-NIR spectrometer equipped with three detector modules (TDM).
Table 1 summarizes the crystallographic parameters of the synthesized samples, including unit cell dimensions, average particle size, dislocation density, bond angles, and bond lengths. As shown in Fig. 2(b), the major diffraction peaks shift toward higher angles upon doping, indicating structural distortion. This shift is associated with a reduction in lattice parameters, unit cell volume, and crystallite size. These changes can be attributed to the smaller ionic radii of Ca2+ (1.00 Å) and Cr3+ (0.615 Å) compared to Bi3+ (1.03 Å) and Fe3+ (0.645 Å).41–43 This observation is consistent with Vegard's law, which states that substituting larger ions with smaller ones generally leads to lattice contraction.44 A decrease in the c-axis with smaller crystallite sizes indicates a contraction along that direction. The merging of the (104) and (110) peaks, as represented by the decreasing values of Δθ (θ110–θ104), may eventually lead to the formation of a single (200) peak in the cubic phase, suggesting a progressive structural transformation upon dopant incorporation.45
| Composition | Unit cell parameter (Å) | Avg. particle size (nm) | Dislocation density ×10−4 (nm−2) | Fe–O bond length (Å) | Fe–O–Fe bond angle (°) |
|---|---|---|---|---|---|
| BFO | a = b = 5.5791, c = 13.868 | 83 | 0.068 | 2.14 | 151.6 |
| BCFO | a = b = 5.5797, c = 13.860 | 74 | 0.130 | 2.05 | 149.2 |
| BCFOCr1 | a = b = 5.5797, c = 13.857 | 71 | 0.172 | 2.05 | 150.7 |
| BCFOCr3 | a = b = 5.5783, c = 13.853 | 67 | 0.205 | 2.04 | 157 |
| BCFOCr5 | a = b = 5.5764, c = 13.852 | 61 | 0.214 | 2.02 | 153 |
Fig. 2(c) and (d) illustrate the relationship between micro-strain, Δθ, c/a ratio, and crystallite size. The c/a ratios of all doped samples are lower than that of pure BFO, which supports the observation of reduced rhombohedral distortion, eventually leading to the formation of a cubic phase.45,46 The average crystallite sizes of the nanoparticles were calculated from XRD peak broadening using the Debye–Scherrer equation. With increasing dopant concentration, both lattice microstrain and dislocation density exhibited an upward trend, reaching their maximum values in the BCFOCr5 sample. The observed variations in microstrain and lattice parameters suggest lattice distortions induced by grain boundaries in nanocrystallites.46,47 In contrast, the crystallite size decreased with higher doping levels, consistent with the average particle size observed in the SEM analysis.
:
Fe ratio. The multi-point EDS analysis for all compositions have been added to the SI (Fig. S1–S5).
![]() | ||
| Fig. 3 SEM image of (a) BFO, (b) BCFO, (c) BCFOCr1, (d) BCFOCr3, and (e) BCFOCr5 with their insets showing representive EDS spectra from a single characteristic point. | ||
Bismuth volatility is a well-documented challenge in bismuth-based oxide systems, including BFO.51,52 In the present study, the undoped BFO sample was found to be significantly bismuth-deficient, as indicated by the Bi
:
Fe ratios in Fig. 3. However, doping with Ca and Cr notably improved the stoichiometry of the perovskite structure, with BCFOCr5 approaching the ideal stoichiometric ratio. Controlled Cr incorporation (1–5%) systematically reduced the non-stoichiometry, as verified by EDS analysis, suggesting that Cr doping helped suppress bismuth volatilization. This improved compositional balance, together with the consistent detection of Ca and Cr in the EDS spectra, supports the successful incorporation of Ca and Cr into the perovskite lattice and likely contributed to the enhanced photocatalytic performance observed in BCFOCr5.
High-resolution TEM (HRTEM) analysis was performed on the BCFOCr5 sample, the composition with the highest photocatalytic activity and most pronounced lattice strain among all the samples, to effectively capture the evolution of lattice strain. Fig. 4(a) and (b) shows the HRTEM image of BCFOCr5 nanoparticles along with the Selected Area Electron Diffraction (SAED) pattern. The polycrystalline nature of the sample, previously confirmed through XRD analysis, is further supported by the SAED pattern [Fig. 4(a)], which shows concentric rings composed of discrete bright spots. The TEM image [Fig. 4(b)] reveals lattice fringes with varying orientations, a characteristic feature of polycrystalline materials.
![]() | ||
| Fig. 4 (a) SAED pattern of BCFOCr5 with indexed planes (b) TEM image and corresponding lattice fringes of BCFOCr5. | ||
Each diffraction ring in the SAED pattern was indexed according to their (hkl) planes by comparing their individual d-spacing values (calculated from ring diameter) against the Powder Diffraction File (PDF) for BiFeO3 (Reference code 96-100-1091). Performing the Fast Fourier Transform (FFT) algorithm on the high-resolution TEM images provided distortion free images of the lattice fringes, depicted in Fig. 4(b) insets, from which inter-planar spacing was calculated. The observed d-spacing values, d1 = 0.28 nm and d2 = 0.39 nm, correspond the (104) and (012) planes of BFO, respectively, using the aforementioned PDF file as reference. These findings are coherent with the results obtained from the SAED pattern. The (012) and (104) planes also provide the highest intensity peaks in the XRD patterns. This, in conjunction with the TEM images, suggest that the NPs exhibit preferential orientation along these crystallographic planes.
Interplanar spacings (dhkl) for selected reflections were calculated from the refined lattice parameters using the hexagonal representation of the rhombohedral R3c structure, according to the relation:
![]() | (1) |
Table 2 compares the interplanar spacings calculated from XRD-refined lattice parameters with those measured directly from HRTEM lattice fringes for BCFOCr5. For BCFOCr5, the interplanar spacings obtained from HRTEM lattice fringes agree closely with values calculated from the refined lattice parameters (0.2810 nm for (104) and 0.3961 nm for (012)). This agreement between diffraction and imaging evidence confirms that the dopant ions are incorporated into the perovskite lattice rather than forming impurity phases. The reduced d-spacings observed in BCFOCr5 by HRTEM compared to BFO (Table 2) are consistent with the lattice contraction and increased microstrain obtained from XRD analysis.
| Composition | Plane (hkl) | dcalchkl (nm) | dTEMhkl (nm) |
|---|---|---|---|
| BFO | (104) | 0.2830 | — |
| BFO | (012) | 0.3962 | — |
| BCFOCr5 | (104) | 0.2810 | 0.2800 |
| BCFOCr5 | (012) | 0.3961 | 0.3900 |
![]() | ||
| Fig. 5 Tauc plot showing direct band gap of (a) BFO, (b) BCFO, (c) BCFOCr1, (d) BCFOCr3, (e) BCFOCr5 (f) schematic illustration of the bandgap narrowing trend. | ||
Oxygen vacancy concentration increases upon doping BFO with Ca. When Ca2+ substitutes Bi3+, a charge imbalance is introduced, leading to the formation of lattice imperfections. The Kröger–Vink notation describing this phenomenon is shown below:
![]() | (2) |
This reaction aligns with the XPS reports in our previous work on the same batch of Ca–Cr co-doped BFO samples, which confirm that the charge imbalance is primarily compensated through the formation of oxygen vacancies, rather than oxidation of Fe3+ to Fe4+.36 The resulting lattice imperfections introduce trap states between the conduction and valence bands, leading to a reduction in the band gap.55 In contrast, Cr3+ doping does not induce additional charge imbalances since Cr and Fe share the same valence state (+3). However, Eg continues to decrease with increasing Cr content. This can be attributed to the interaction between conduction band electrons and the localized d-electrons of Cr ions at Fe sites, which leads to band gap narrowing through sp–d spin exchange interactions.28 Additionally, Cr doping introduces defect-induced energy levels near the conduction band, which contribute to further band gap reduction, particularly in smaller particles where surface defect densities are higher.45 Structural distortion of the FeO6 octahedra caused by doping alters Fe–O–Fe bond angles and lengths, rearranging molecular orbitals in a way that modifies the electronic band structure and effectively narrows the band gap.56 Lastly, the microstrain arising from non-uniform lattice distortions, dislocations, and grain boundary relaxation plays a significant role in modifying local energy levels and shifting the absorption edge.45
The concentration of MB in the solution was monitored before and after photodegradation by extracting aliquots at fixed time intervals and analyzing them using a UV-Vis spectrophotometer. The absorption spectra of the MB dye solution at various time points in the presence of suspended nanoparticles are shown in Fig. 6(a–e). Upon exposure to the xenon lamp, the characteristic blue color of the MB solution gradually faded, indicating progressive degradation of the dye and a corresponding shift in the absorption peaks.
Degradation kinetics were evaluated using a plot of
versus time, in accordance with the Langmuir–Hinshelwood model, confirming that the reaction follows pseudo-first-order kinetics
![]() | (3) |
![]() | (4) |
After 90 minutes of illumination, the degradation efficiencies were found to be 71%, 72%, 76%, 78%, and 93% for BFO, BCFO, BCFOCr1, BCFOCr3, and BCFOCr5, respectively [Fig. 6(h)], indicating a significant improvement in the photocatalytic performance with increasing dopant concentration. It is worth noting that the 93% degradation efficiency in BCFOCr5 observed here is among the highest reported for BFO systems explicitly tested at neutral pH.
The corresponding reaction rate constants (k) were 0.01358, 0.01450, 0.01586, 0.01700, and 0.03038 min−1 for BFO, BCFO, BCFOCr1, BCFOCr3, and BCFOCr5, respectively [Table 3], reconfirming the critical role of dopant concentration in modulating photocatalytic activity. The systematically enhanced rate constants observed here are consistent with improved charge separation upon Ca–Cr doping. Although photoluminescence (PL) measurements would directly probe carrier recombination, our photocatalytic kinetics already provide clear evidence of suppressed recombination. This enhancement stems from the combined effects of band gap narrowing and particle size reduction induced by Ca and Cr co-doping. A narrower Eg enhances visible light absorption and boosts the generation of electron–hole pairs, while smaller particle sizes increase surface area, offering more active sites for redox reactions.58,59 Doping also creates trap states that hinder charge carrier recombination, thereby improving overall efficiency. Additionally, at higher dopant levels, the formation of a core–shell structure further strengthens photocatalytic activity by establishing a localized surface electron cloud that functions as an internal electric field, promoting effective charge separation, as discussed in Section 3.6.
| Composition | Eg (eV) | Dye | k (min−1) |
|---|---|---|---|
| BFO | 2.13 | MB | 0.01358 |
| BCFO | 2.07 | 0.01450 | |
| BCFOCr1 | 2.06 | 0.01586 | |
| BCFOCr3 | 2.01 | 0.01700 | |
| BCFOCr5 | 1.80 | 0.03038 |
Under xenon lamp irradiation, the Bi0.97Ca0.03Fe1−xCrxO3 nanoparticles absorb photons with energies equal to or greater than its Eg [eqn (5)], leading to the generation of electron–hole pairs. These charge carriers migrate to the catalyst surface, where they participate in redox reactions that produce Reactive Oxygen Species (ROS), primarily superoxide radicals (O2˙−) via oxygen reduction [eqn (6)] and hydroxyl radicals (˙OH) through water oxidation [eqn (7)]. The resulting ROS actively degrade dye molecules, as shown in eqn (8) and (9). Additionally, direct degradation by photogenerated holes is depicted in eqn (10).58,59 These reactive radicals oxidize MB into carbon dioxide (CO2) and water (H2O), while converting nitrogen (N) and sulfur (S) heteroatoms into inorganic byproducts like nitrate, ammonium, and sulfate ions, respectively.60 A schematic overview of this mechanism is presented in Fig. 7.
| BCFOCr + hυ = e− + h+ | (5) |
| O2 + e− = O2˙− | (6) |
| H2O + h+ = ˙OH | (7) |
| O2˙− + MB = CO2 + H2O + Cl− + SO42− + NO3− | (8) |
| ˙OH + MB = CO2 + H2O + Cl− + SO42− + NO3− | (9) |
| h+ + MB = CO2,+ H2O,+ Cl− + SO42− + NO3− | (10) |
Table 4 presents a comparative analysis of pure and doped BFO systems, summarizing their synthesis methods, Eg, pH conditions, and photocatalytic efficiencies. As seen, the co-doped BCFOCr5 sample exhibits the highest degradation efficiency (93%) under explicitly neutral pH. This performance was achieved solely through compositional tuning, without any external pH adjustment. Since MB degradation tends to accelerate under alkaline conditions, further optimization of pH could potentially yield complete dye removal.
| Composition | Synthesis method | Eg (eV) | Dye | pH | Degradation efficiency | Ref. |
|---|---|---|---|---|---|---|
| BFO | Sol–gel | 2.36 | MB | 7 | 69% in 240 min | 64 |
| BFO | Hydrothermal | 2.12 | MB | 7–8 | 52% in 180 min | 65 |
| BFO | Sol–gel | 2.02 | MB | 7 | 81% in 150 min | 24 |
| BFO | Sol–gel | 2.02 | MB | 3 | 99% in 90 min | 24 |
| BFO | Hydrothermal | 2.08 | MB | 6.8 | 45% in 90 min | 66 |
| BFO | Sono-synthesis | 2.17 | MB | 6 | 10% in 90 min | 26 |
| BFO | Sono-synthesis | 2.17 | MB | 2.5 | 100% in 80 min | 26 |
| BFO | Sol–gel | 1.91 | MB | 6.7 | 100% in 150 min | 23 |
| BFO | Sol–gel | 1.91 | MB | 11 | 100% in 30 min | 23 |
| BFO | Biotemplate | Not reported | MB | 8–10 | 96% in 180 min | 67 |
| BFO | Sol–gel | 1.8 | MB | 8 | 98.49% in 120 min | 68 |
| BFO | Solgel | 1.8 | MB | 4.3 | 96.2% in 180 min | 68 |
| BFO | Sol–gel | 2.38 | MB | Not reported | 58% in 240 min | 22 |
| Bi0.9Gd0.1FeO3 | Sol–gel | 2.32 | MB | Not reported | 94% in 240 min | 22 |
| Bi0.95Sr0.05Fe0.99Cr0.01O3 | Sol–gel | 1.86 | MB | Not reported | 92.9% in 90 min | 31 |
| Bi0.95Sm0.05Fe0.8Mn0.2O3 | Sol–gel | 1.52 | MB | Not reported | 65% in 180 min | 28 |
| Bi0.85Sm0.15Fe0.95Cr0.05O3 | Sol–gel | 1.96 | MB | Not reported | 99.16% in 90 min | 29 |
| Bi0.90Gd0.05Sm0.05FeO3 | Solvo-thermal | 1.96 | MB | 2.24 | 95% in 40 min | 30 |
| Bi0.95Y0.05FeO3 | Biotemplate | 2.05 | MB | 8 | 94.5% in 180 min | 69 |
| Bi0.985Sn0.015FeO3 | Sol–gel | 1.94 | MB | Not reported | 99% in 120 min | 70 |
| Bi0.9Gd0.1FeO3 | Hydrothermal | 1.18 | MB | Not reported | 97% in 180 min | 71 |
| Co doped BFO | Polyol | Not reported | MB | 3.5 | 35.06% in 120 min | 72 |
| BCFOCr5 | Sol–gel | 1.77 | MB | 7 | 93% in 90 min | Present work |
Furthermore, the increased saturation magnetization significantly improves the potential for magnetic separation, which is a key factor in enabling the recovery and reuse of photocatalysts for cost-effective, sustainable, and scalable applications.28,61–63 Previous studies have already demonstrated the reusability of BFO catalysts. For example, Soltani et al. reported that pure BFO retained its degradation efficiency in rhodamine B dye without any loss of activity across four cycles,25 while Jhansi et al. observed unchanged structural, morphological, and dye degradation properties of BFO nanoparticles across three cycles in MB.23 Based on these findings, it is reasonable to anticipate that the magnetically recoverable Ca–Cr co-substituted BFO nanoparticles explored in this study hold strong potential for practical reuse in wastewater treatment applications,73 even though a detailed recycling study lies beyond the scope of the present work.
Supplementary information: multi-point EDS spectra for all compositions (BFO, BCFO, and Ca–Cr co-doped BiFeO3 variants) used in this study. See DOI: https://doi.org/10.1039/d5ra06225j.
Footnote |
| † Anika Rahman Riya, Abrar Daiyan, Mehedi Hasan Prince, Troyee Mitra Aishi contributed equally to this work. |
| This journal is © The Royal Society of Chemistry 2025 |