Next Article in Journal
Enhancing Wireless Power Transfer Performance Based on a Digital Honeycomb Metamaterial Structure for Multiple Charging Locations
Previous Article in Journal
Modulation of Surface Elastic Waves and Surface Acoustic Waves by Acoustic–Elastic Metamaterials
Previous Article in Special Issue
Tailoring the Synthesis Method of Metal Oxide Nanoparticles for Desired Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cr3+ Doping Effects on Structural, Optical, and Morphological Characteristics of BaTiO3 Nanoparticles and Their Bioactive Behavior

by
Efracio Mamani Flores
1,
Bertha Silvana Vera Barrios
2,*,
Julio César Huillca Huillca
1,
Jesús Alfredo Chacaltana García
1,
Carlos Armando Polo Bravo
1,
Henry Edgardo Nina Mendoza
1,
Alberto Bacilio Quispe Cohaila
3,4,
Francisco Gamarra Gómez
5,
Rocío María Tamayo Calderón
6,
Gabriela de Lourdes Fora Quispe
4,7 and
Elisban Juani Sacari Sacari
4,8,*
1
Departamento de Física Aplicada, Facultad de Ciencias, Universidad Nacional Jorge Basadre Grohmann, Avenida Miraflores S/N, Ciudad Universitaria, Tacna 23003, Peru
2
Facultad Ing. de Minas, Universidad Nacional de Moquegua, Moquegua 18001, Peru
3
Laboratorio de Generación y Almacenamiento de Hidrogeno, Facultad de Ingeniería, Universidad Nacional Jorge Basadre Grohmann, Avenida Miraflores S/N, Ciudad Universitaria, Tacna 23003, Peru
4
Grupo de Investigación GIMAECC, Facultad de Ingeniería, Universidad Nacional Jorge Basadre Grohmann, Avenida Miraflores S/N, Ciudad Universitaria, Tacna 23003, Peru
5
Laboratorio de Nanotecnología, Facultad de Ingeniería, Universidad Nacional Jorge Basadre Grohmann, Avenida Miraflores S/N, Ciudad Universitaria, Tacna 23003, Peru
6
Centro de Microscopia Electrónica, Facultad de Ingeniería de Procesos, Universidad Nacional de San Agustín, Arequipa 04001, Peru
7
Laboratorio de Biorremediación, Facultad de Ciencias, Universidad Nacional Jorge Basadre Grohmann, Avenida Miraflores S/N, Ciudad Universitaria, Tacna 23003, Peru
8
Facultad de Ciencias, Universidad Nacional de Ingeniería, Av. Túpac Amaru 210, Lima 15333, Peru
*
Authors to whom correspondence should be addressed.
Crystals 2024, 14(11), 998; https://doi.org/10.3390/cryst14110998 (registering DOI)
Submission received: 9 August 2024 / Revised: 16 November 2024 / Accepted: 17 November 2024 / Published: 19 November 2024
(This article belongs to the Special Issue Synthesis and Characterization of Oxide Nanoparticles)

Abstract

:
This study investigates the effects of chromium (Cr3+) doping on BaTiO3 nanoparticles synthesized via the sol–gel route. X-ray diffraction confirms a Cr-induced cubic-to-tetragonal phase transition, with lattice parameters and crystallite size varying systematically with Cr3+ content. UV–visible spectroscopy reveals a monotonic decrease in bandgap energy from 3.168 eV (pure BaTiO3) to 2.604 eV (5% Cr3+-doped BaTiO3). Raman and FTIR spectroscopy elucidate structural distortions and vibrational mode alterations caused by Cr3+ incorporation. Transmission electron microscopy and energy-dispersive X-ray spectroscopy verify nanoscale morphology and successful Cr3+ doping (up to 1.64 atom%). Antioxidant activity, evaluated using the DPPH assay, shows stable radical scavenging for pure BaTiO3 (40.70–43.33%), with decreased activity at higher Cr3+ doping levels. Antibacterial efficacy against Escherichia coli peaks at 0.5% Cr3+ doping (10.569 mm inhibition zone at 1.5 mg/mL), decreasing at higher concentrations. This study demonstrates the tunability of structural, optical, and bioactive properties in Cr3+-doped BaTiO3 nanoparticles, highlighting their potential as multifunctional materials for electronics, photocatalysis, and biomedical applications.

1. Introduction

Barium titanate (BaTiO3) nanoparticles have garnered significant attention in materials science due to their exceptional ferroelectric, piezoelectric, and dielectric properties [1,2]. These nanoparticles, belonging to the perovskite family with the general formula ABO3 [3], exhibit a unique crystal structure that undergoes several phase transitions, each associated with distinct functional characteristics. At the nanoscale, BaTiO3 exhibits distinct advantages over its bulk counterpart. Nanoparticles with diameters below 100 nm show a reduced Curie temperature of 30–50 °C compared to bulk, extending their ferroelectric behavior to lower temperatures [4,5]. The dielectric constant increases significantly, with values up to 15,000 reported for 70 nm particles, enhancing capacitor performance [6]. Piezoelectric coefficients are also amplified, with a d33 of 416 pC/N observed in 100 nm BaTiO3, compared to 216 pC/N in 500 nm particles [7]. Additionally, the surface area can increase up to seven times with the reduction of particle size from 70 nm to 10 nm, which can enhance the catalytic activity [8]. These quantifiable nanoscale-specific properties make BaTiO3 nanoparticles promising for applications in miniaturized electronics, high-sensitivity sensors, and efficient photocatalysts [2].
Various synthesis methods have been employed to produce BaTiO3 nanoparticles, each offering distinct advantages and challenges. Conventional solid-state reactions, while straightforward, often require high temperatures and yield particles with limited size control [9]. Coprecipitation methods allow for lower processing temperatures but can struggle with precise stoichiometry control [10]. Hydrothermal synthesis offers good control over particle morphology and size, but is limited by small production volumes due to high-pressure autoclave reactors [11]. Other methods such as microwave-assisted and combustion synthesis have also been investigated [12,13]. Among these techniques, the sol–gel method stands out for its versatility and effectiveness in producing high-quality BaTiO3 nanoparticles [14], particularly when incorporating dopants. This method involves the transition of a system from a colloidal suspension (sol) into a three-dimensional network structure that entraps the solvent (gel), allowing for molecular-level mixing of the precursors and dopants. The gel state, an intermediate phase in this process, plays a crucial role in controlling the evolution of the material, enabling precise tailoring of the composition, particle size, and morphology [15]. Furthermore, the sol–gel approach facilitates homogeneous dopant distribution and allows for the synthesis of nanocrystalline materials at relatively low temperatures [16], making it particularly suitable for studying the effects of dopant incorporation on BaTiO3 properties.
The incorporation of dopants into the BaTiO3 lattice has emerged as a powerful strategy to further tailor its properties [17]. Among various dopants, trivalent chromium (Cr3+) has shown particular promise due to its ability to simultaneously modulate the structural, optical, and electrical properties of BaTiO3 nanoparticles. The ionic radius of Cr3+ (0.69 Å) closely matches that of Ti4+ (0.68 Å), facilitating its substitution at the B-site of the perovskite structure [18]. This substitution introduces complex defect chemistry, profoundly affecting the characteristics of the material at multiple scales.
From a structural perspective, Cr3+ doping induces lattice distortions and alters the phase transition behavior of BaTiO3 nanoparticles [19]. These structural modifications can significantly influence the material’s ferroelectric and piezoelectric responses, potentially enhancing its performance in various applications [20]. The precise control over the crystal structure through Cr3+ doping offers a unique opportunity to tailor the functional properties of BaTiO3 nanoparticles for specific technological needs.
The optical properties of BaTiO3 nanoparticles undergo substantial changes upon Cr3+ doping. The introduction of Cr3+ ions creates additional electronic states within the bandgap, leading to a reduction in the optical bandgap energy [21]. This bandgap engineering not only alters the optical absorption characteristics of the material, resulting in visible color changes, but also expands its potential for photocatalytic and optoelectronic applications [22]. The ability to tune the optical absorption characteristics through Cr3+ doping opens up new avenues for utilizing BaTiO3 nanoparticles in light-driven technologies.
Morphologically, Cr3+ doping can significantly influence the size, shape, and surface properties of BaTiO3 nanoparticles [23,24]. These morphological changes are crucial as they directly impact the material’s reactivity, catalytic activity, and interaction with its environment. Understanding and controlling the morphological evolution of Cr3+-doped BaTiO3 nanoparticles is essential for optimizing their performance in various applications, particularly those involving interface-dependent processes [25].
Perhaps most intriguingly, recent research has hinted at the potential bioactive behavior of Cr3+-doped BaTiO3 nanoparticles, opening up exciting possibilities in the realm of biomedical applications [3]. The unique combination of ferroelectric properties and controlled doping may impart novel functionalities relevant to biological interactions [26]. These could include enhanced biocompatibility, antimicrobial activity, or the ability to influence cellular processes through localized charge distributions or piezoelectric effects [27].
This study aims to provide a comprehensive investigation of the effects of Cr3+ doping on the structural, optical, and morphological characteristics of BaTiO3 nanoparticles, with a particular focus on their resultant bioactive behavior. Using advanced characterization techniques, we aim to unravel the complex relationships between doping concentration, nanoparticle characteristics, and biological interactions. This research contributes to the fundamental understanding of doped ferroelectric nanomaterials but also paves the way for their rational design in next-generation applications in biomedicine.

2. Materials and Methods

2.1. Materials

High-purity analytical grade materials were used for the preparation of pure and Cr3+-doped barium titanate, including barium nitrate (Ba[NO3]2, Merck, 99.0%), titanium (IV) isopropoxide (C12H28O4Ti, Sigma Aldrich, 97%, St. Louis, MO, USA), chromium (III) nitrate nonahydrate (CrN3O9.9H2O, Sigma Aldrich, 99%, MO, USA), citric acid (C6H8O7, Sigma Aldrich, MO, USA), 2-propanol (CH3CH[OH]CH3, Merck, Darmstadt, Germany), DPPH (2,2-Difenil-1-(2,4,6-trinitrofenil)hidrazil, Merck Darmstadt, Germany), Mueller–Hinton agar (Merck Darmstadt, Germany), and ultrapure water.

2.2. Pure and Cr3+-Doped BaTiO3 Synthesis

The sol–gel method was employed for synthesizing pure and Cr3+-doped BaTiO3 nanoparticles. This method was chosen for its ability to achieve molecular-level mixing of the precursors and dopants, facilitating homogeneous composition and precise control over the doping process.
Ultrapure water (20 mL) was used to dissolve barium nitrate (6.25 mmol), while 2-propanol served in two steps: 5 mL to dissolve citric acid (1.7 mmol) and another 5 mL for titanium isopropoxide (6.25 mmol). For Cr3+-doping, barium nitrate was partially substituted with chromium nitrate (0.03, 0.05, 0.1, 0.3, and 0.5 mol%). These solutions were prepared separately and combined sequentially, with the barium nitrate and citric acid solutions mixed and stirred for 150 min before the dropwise addition of the titanium isopropoxide solution. The gel state, characterized by a three-dimensional network structure entrapping the solvent, is crucial in this process. It allows for the homogeneous distribution of dopants at the molecular level, particularly important for achieving uniform Cr3+ incorporation into the BaTiO3 lattice. This stage significantly influences the final properties of the material.
Gels underwent aging at 50 °C for 12 h, during which further condensation reactions and structural reorganization occurred. The dried gels were ground into fine powders and thermally treated at 700 °C for two hours, causing gel network breakdown, organic component removal, and BaTiO3 phase crystallization.

2.3. Characterization

A comprehensive array of analytical techniques was utilized to characterize the structural, compositional, thermal, optical, vibrational, and morphological properties of pure and chromium-doped barium titanate (BaTiO3) ceramic samples. Simultaneous thermogravimetric analysis and differential scanning calorimetry were conducted on the dried gel precursors prior to calcination using an SDT 650 module from TA Instruments (New Castle, DE, USA) to study the thermal decomposition behavior of the precursors, determine the optimal calcination temperature for crystalline BaTiO3 formation, and quantify changes in thermal behavior and weight loss. The samples were heated from room temperature to 900 °C at a constant rate of 20 °C per minute under a nitrogen atmosphere. X-ray diffraction patterns were collected using a PANalytical Aeris diffractometer (PANalytical, Almelo, The Netherlands) with copper K-alpha radiation and a nickel filter, and Rietveld refinement of the patterns enabled the determination of the crystal structure and crystallite size. Additionally, vibrational spectroscopy based on a 532 nm Raman microscope from Ocean Optics (MAYA 2000-Pro, Ocean Optics, Dunedin, FL, USA) was utilized to acquire Raman spectra. Diffuse reflectance spectroscopy in the ultraviolet–visible wavelength range was performed using an Evolution 220 spectrometer from Thermo Scientific (San Jose, CA, USA). Infrared spectroscopy was performed on a Bruker Invenio R Fourier Transform Infrared spectrometer equipped with an attenuated total reflectance accessory (Bruker, Saarbrucken, Germany). Photoluminescence spectra were recorded using a Fluoromax Plus spectrometer (Horiba Scientific, Irvine, CA, USA). The morphology and microstructure were evaluated using transmission electron microscopy on a Thermo Scientific Talos F200i microscope (Thermo Scientific Co., Eindhoven, The Netherlands).

2.4. Antioxidant Activity

Pure and doped Barium titanate nanoparticle solutions were prepared at varying concentrations ranging from 0.5 to 4 mg/mL. The radical scavenging activity was assessed by reacting the BaTiO3 nanoparticles solutions with the stable free radical 2,2-diphenyl-1-picrylhydrazyl (DPPH). Specifically, 50 μL of each sample concentration was added in triplicate to the wells of a 96-well microplate followed by the addition of a 50 μL methanolic solution of 100 μM DPPH. Control wells contained 50 μL DPPH solution and 50 μL sterile deionized water. The microplate was incubated at room temperature in the dark for 20 min. Thereafter, the absorbance was measured at 517 nm using a spectrophotometer (Epoch 2C, Biotek Instruments, Winooski, VT, USA). The assays were performed in three independent replicates. Radical scavenging percentages were computed using the following equation:
Percent   scavenging   ( % ) = A 0 A 1 A 0 100
where A0 = absorbance of DPPH control and A1 = absorbance of DPPH solution after reaction with the pure and doped BaTiO3 nanoparticles.

2.5. Antibacterial Activity Assay

The antimicrobial activity of pure and doped BaTiO3 nanoparticles was evaluated against E. coli using the Kirby–Bauer disk diffusion method. E. coli was cultured in Mueller–Hinton broth at 37 °C for 24 h prior to use on agar plates. E. coli inoculum was prepared by adjusting the turbidity to match the 0.5 McFarland standard, corresponding to an approximate cell density of 1.5 × 108 CFU/mL. Mueller–Hinton agar plates were inoculated with 100 μL of bacterial suspension evenly spread across the surface. Sterile filter paper discs (6 mm diameter) (Oxoid Ltd., Basingstoke, UK) were prepared from deactivated sensitivity test discs through a validated sterilization protocol (autoclaving at 121 °C for 15 min in distilled water, followed by triple rinsing with sterile distilled water and dry heat sterilization at 180 °C for 1 h). Complete deactivation was verified through control experiments. These blank discs were then impregnated with varying concentrations of pure and doped BaTiO3 nanoparticle suspensions (0.5–4 mg/mL). The impregnated discs were dried at room temperature under sterile conditions before being placed on the inoculated agar plates. BaTiO3 nanoparticle solutions were prepared in sterile deionized water. The agar plates were incubated at 37 °C for 24 h under aerobic conditions. Deionized water was used as a negative control in parallel with all antibacterial experiments. The negative control experiments were conducted under identical conditions to those used for testing the Cr3+-doped BaTiO3 nanoparticles.
Finally, the diameters of the inhibition zones formed around the disks were measured to evaluate the antimicrobial activity. The assays were performed in triplicate and the mean inhibition zone diameters were recorded for each BaTiO3 nanoparticle concentration.

3. Results and Discussion

3.1. Thermal Analysis

Thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) were employed to investigate the thermal behavior of pure and Cr3+-doped BaTiO3 nanoparticles synthesized using citric acid as the chelating agent (Figure 1). The TGA curves (Figure 1a) reveal varied weight loss patterns across the doping concentrations. Notably, the 5% Cr3+-doped sample exhibits a higher mass loss compared to the undoped sample, suggesting that Cr3+ doping has a complex effect on the thermal behavior of BaTiO3 nanoparticles. This may involve influencing the retention of volatile species or altering decomposition pathways of residual organics, particularly the citric acid chelate complexes [28,29].
The DSC curves (Figure 1b) provide insight into the energetic processes occurring during heating. An endothermic event around 100 °C is primarily attributed to the evaporation of residual moisture. The subsequent thermal events in the 200–500 °C range correspond to a complex series of decomposition processes, primarily involving the citric acid used as the chelating agent in our synthesis [30]. Citric acid decomposition typically occurs over a broad temperature range (177–500 °C) [31], which aligns with our observations. The exothermic nature of these events suggests the combustion of organic material, releasing energy that may facilitate the initial stages of BaTiO3 formation.
The exothermic peak observed around 600 °C corresponds to the crystallization of BaTiO3. Interestingly, the peak temperatures for Cr3+-doped samples are slightly higher than the undoped sample, suggesting that Cr3+ doping may increase the energy required for crystallization. This could be due to lattice distortions, altered nucleation dynamics, or the influence of Cr3+ on the decomposition products of the citric acid complexes [28,32].
The role of Cr3+ in the crystallization and phase transition of BaTiO3 is complex. While Cr3+ doping appears to slightly increase the crystallization temperature, it may still influence the subsequent cubic-to-tetragonal transition. Recent studies have shown that transition metal dopants can create local distortions in the BaTiO3 lattice, potentially serving as nucleation points for the tetragonal phase [33,34,35]. However, this transition is not distinctly visible in our DSC data, likely due to the small energy difference between the cubic and tetragonal phases in nanocrystalline BaTiO3. The exact mechanism in our Cr3+-doped samples requires further investigation, potentially using in situ diffraction techniques to resolve the subtle structural changes during heating.
The varied weight loss patterns observed in TGA, particularly for higher Cr3+ doping levels, may be attributed to the formation of different metal–citrate complexes during synthesis. These complexes could have different thermal stabilities and decomposition pathways, leading to the observed variations in weight loss. This complex interplay between Cr3+ doping, citric acid complexation, and thermal decomposition underscores the intricate nature of the synthesis–structure–property relationships in these doped nanoparticles.

3.2. X-Ray Diffraction

The X-ray diffraction (XRD) patterns are shown in Figure 2a. The analysis reveals the influence of chromium (Cr3+) doping on the structural transformation of BaTiO3, where Cr3+ was introduced through chromium nitrate dissolved in 2-propanol during the sol–gel synthesis. The pristine BaTiO3 exhibits a cubic crystal system (Pm-3m space group), characterized by a single peak in the XRD pattern at 2θ ≈ 45.75° (Figure 2b). The observed XRD pattern for the cubic phase shows good agreement with the reference pattern (Figure S1) from the Crystallography Open Database [36] (COD 96-591-0150). Upon Cr3+ doping at the A site through the controlled substitution of Ba2+ during the sol–gel process, a structural transition to the tetragonal phase (P4mm space group) is observed (Figure 2b), evidenced by the splitting of the original peak into a doublet, with a new shoulder appearing at 2θ ≈ 45.25° for Cr3+-doped samples. The XRD pattern of the Cr3+-doped samples matches well with the COD entry 96-152-5438, indicating the successful formation of the tetragonal phase. These findings are consistent with the well-established concept of dopant-induced phase transitions in perovskite oxides [37,38].
The effectiveness of using chromium nitrate dissolved in 2-propanol for Cr3+ incorporation is demonstrated by the systematic evolution of the XRD patterns with increasing dopant concentration. The homogeneous distribution of Cr3+ ions achieved through the sol–gel route enables uniform substitution at the A-site, as evidenced by the consistent peak splitting patterns across different doping levels.
The minor peaks observed around 24.37°, 24.78°, 34.83°, 42.45°, 43.46°, and 46.9° can be attributed to trace amounts of BaCO3 (COD: 96-900-6842), formed by the reaction of Ba2+ with atmospheric CO2 during synthesis. These peaks correspond to the (111), (102), (013), (203), (104), and (114) planes of BaCO3, respectively.
The sol–gel synthesis approach using chromium nitrate in 2-propanol plays a crucial role in achieving the observed structural characteristics, where molecular-level mixing in the sol state and controlled hydrolysis and condensation reactions enable precise control over the Cr3+ substitution process. This is particularly important given the significant difference in ionic radii between Ba2+ (1.35 Å) and Cr3+ (0.69 Å). Furthermore, the sol–gel method has demonstrated versatility in accommodating various dopants, such as calcium [39], lanthanum [40], and cerium [41], among others [17,42], which have smaller ionic radii compared to Ba2+.
The Rietveld refinement analysis (Figure S2) reveals systematic changes in structural parameters with increasing Cr3+ content, as presented in Table 1. The pristine BaTiO3 sample exhibits equal lattice parameters (a = b = c = 4.00785 Å) and 90° angles between them, characteristic of the cubic phase. Upon Cr3+ doping, the structural transition to the tetragonal symmetry, with the a = b lattice parameter decreasing from 4.00698 Å to 4.00379 Å, and the c parameter increasing from 4.01477 Å to 4.01699 Å (Figure S3) [43,44].
The defect chemistry associated with this A-site substitution can be described using Kröger–Vink notation:
2 C r ( N O 3 ) 3 + 3 B a B a × 2 C r B a + V B a + 3 B a   ( N O 3 ) 2
C r ( N O 3 ) 3 + B a B a × C r B a + O i + B a   ( N O 3 ) 2
where C r B a represents Cr3+ on a Ba2+ site (single positive charge), V B a represents a barium vacancy (double negative charge), and O i represents an oxygen interstitial (double negative charge).
The crystallite size analysis shows a marked decrease from 63.27 nm (pristine) to 55.29 nm (0.3% Cr3+), followed by relatively stable values around 54–58 nm for higher doping levels. The micro strain analysis from the Rietveld refinement shows values decreasing from 0.097% (pristine) to 0.046% (5% Cr3+), indicating that despite the significant size mismatch between Ba2+ and Cr3+, the sol–gel synthesis conditions enabled the formation of a relatively stable crystal structure [38].

3.3. Raman Spectroscopy

The Raman spectra (Figure 3) provide detailed insights into the local structural changes induced by A-site Cr3+ substitution in BaTiO3. The pristine BaTiO3 (Figure 3a) exhibits characteristic first-order vibrational modes with F2g symmetry, representing symmetric Ti-O bending (305 cm−1) and stretching (720 cm−1) motions. These modes reflect the cubic symmetry of the undoped structure, where TiO6 octahedra maintain regular coordination geometry [45].
Upon Cr3 incorporation at the A-site (Figure 3b), systematic changes in the vibrational spectra emerge, reflecting the modified crystal structure. The development of additional modes at 153, 521, 637, and 719 cm−1 indicates increasing lattice distortion with Cr3 content. These spectral changes can be attributed to the formation of defect complexes ( C r B a , V B a , and O i ) that modify local symmetry, strain fields created by the significant size mismatch between Cr3+ and Ba2+, and enhanced displacement of Ti4+ ions from their central positions in response to the modified A-site environment [46].
The systematic evolution of peak intensities and the emergence of new vibrational features correlate with the increasing tetragonal distortion observed in XRD analysis. Notably, samples with higher Cr3+ content (1%, 3%, and 5%) show more pronounced spectral changes, consistent with greater structural modification at elevated doping levels. The low-frequency band at 153 cm−1 can be assigned to A-site cation motions, while the additional features in the 500–700 cm−1 range reflect modified Ti-O bonding environments resulting from the accommodation of Cr3+ in the lattice [18,21,47]. These spectroscopic observations provide compelling evidence for the successful incorporation of Cr3+ at the A-site and the consequent local structural modifications, despite the significant ionic radius mismatch. The Raman data thus support our proposed defect compensation mechanism and provides valuable insights into the local structural changes that enable stable A-site substitution in this system [48].

3.4. Fourier-Transform Infrared Spectroscopy (FTIR)

FTIR spectra (Figure 4) reveal systematic modifications in the vibrational characteristics of BaTiO3 upon A-site Cr3⁺ substitution. The spectrum of pristine BaTiO3 shows characteristic absorption bands at 507 cm−1 (Ti-O bending), 632 cm−1 (asymmetric O-Ti-O stretching), 857 cm−1 (asymmetric Ti-O-Ti bridging stretch), and 1417 cm−1 (Ti-O stretch), consistent with previous reports [49,50]. Upon Cr3⁺ incorporation, several systematic changes emerge in the spectral features. The relative intensity of the 507 cm−1 band decreases with increasing Cr3⁺ content, while new features develop in the 600–700 cm−1 region. These changes reflect modifications in the Ti-O bonding environments induced by the accommodation of Cr3⁺ at the A-site [51].
The band at 695 cm−1, associated with symmetric O-Ti-O stretching in TiO₆ octahedra, shows progressive modification in intensity and shape with increasing Cr3⁺ concentration [48]. This evolution indicates that A-site substitution influences the local symmetry of the octahedral units, likely through the formation of defect complexes involving barium vacancies and oxygen interstitials. The high-frequency mode at 1417 cm−1, corresponding to localized oxygen vibrations, exhibits systematic changes in intensity that correlate with the increasing Cr3⁺ content [52]. Additional weak features observed in samples with higher Cr3⁺ concentrations may indicate the presence of trace secondary phases, although these remain below the detection limit of XRD [53].
The systematic evolution of the FTIR spectra provides further evidence for the successful incorporation of Cr3⁺ at the A-site and the resulting modifications in the local bonding environments. The observed spectral changes align with our proposed defect compensation mechanism, where the significant size mismatch between Cr3⁺ and Ba2⁺ is accommodated through the formation of ordered defect structures. These spectroscopic signatures complement our XRD and Raman analyses, offering additional insights into the local structural modifications that enable stable A-site substitution in this system.

3.5. UV–Visible Spectroscopy

The UV–vis diffuse reflectance spectra, as shown in Figure 5a, reveal the influence of Cr3+ doping on the optical properties of BaTiO3. Pure BaTiO3 exhibits a sharp absorption edge and low reflectance below 375 nm, indicating its wide bandgap and strong UV absorption. As the Cr3+ doping concentration increases, the reflectance progressively decreases across the visible range, correlating with a visible color change from white for undoped BaTiO3 to pale grey hues in the doped ceramics. This qualitative observation suggests a narrowing of the bandgap due to Cr3+ incorporation.
To quantitatively evaluate the optical bandgap values, the Kubelka–Munk function, which relates the absorption coefficient to diffuse reflectance, was employed. The Kubelka–Munk function, F(R), is defined as:
F R = K S = 1 R 2 2 R
where K is the absorption coefficient of radiation, S is the scattering factor, and R is the ratio of the intensities of radiation reflected in a diffuse manner. This function assumes that the transitions occur between extended states, consistent with the electronic structure of BaTiO3.
The calculated bandgap values demonstrate a monotonic decline from 3.168 eV for pure BaTiO3 to 3.03 eV (0.3% Cr3+), 2.939 eV (0.5% Cr3+), 2.834 eV (1% Cr3+), 2.775 eV (3% Cr3+), and finally to 2.604 eV in the 5% Cr3+-substituted composition, validating the significant bandgap engineering achieved through controlled Cr3+ dopant inclusion.
The slightly enhanced bandgap reduction observed at 1% Cr3⁺ doping suggests an optimal configuration of defect complexes at this concentration, leading to more effective modification of the electronic band structure.
This quantifies the magnitude of bandgap engineering achieved through controlled dopant inclusion. The decrease in bandgap energy with increasing Cr3+ concentration can be attributed to the increased hybridization between the Cr 3d and O 2p orbitals localized within the distorted TiO6 octahedral units [54,55]. The structural transitions induced by Cr3+ doping, as evidenced by XRD and Raman spectroscopy, enhance the bonding covalence, facilitating the reduction in bandgap energy.
Moreover, the polarization-strain effects associated with the tetragonal distortion in Cr3+-doped BaTiO3 can assist in charge carrier separation and mobility, further contributing to the observed optical properties. The introduction of Cr3+ dopants also creates mid-gap states, which can act as intermediate energy levels for sub-bandgap optical absorption [56]. These mid-gap states allow for the absorption of photons with energies lower than the intrinsic bandgap of BaTiO3, effectively extending the absorption range into the visible region.

3.6. Photoluminescence Spectroscopy

The photoluminescence (PL) spectra of pure and Cr3⁺-doped BaTiO3 samples (Figure 6), recorded under carefully controlled experimental conditions (excitation: 254 nm, temperature: 25 °C, and fixed illumination area), reveal complex emission behavior reflecting the electronic structure modifications induced by A-site substitution. Gaussian deconvolution analysis (Figure S4, R2 > 0.999) demonstrates the presence of multiple emission components, indicating diverse radiative recombination pathways in the material.
The pristine BaTiO3 exhibits three primary emission components centered at 361.17, 396.74, and 415.92 nm, with relative contributions of 7.08%, 41.17%, and 51.75%, respectively (Table S1). The dominant peak at 415.92 nm (FWHM = 47.69 nm) originates from radiative recombination involving oxygen vacancy states, while the bands at 361.17 nm (FWHM = 60.64 nm) and 396.74 nm (FWHM = 32.77 nm) correspond to near-band-edge transitions and shallow defect states [57], respectively. Upon Cr3⁺ incorporation, the emergence of a fourth emission component (~430–433 nm) is observed in all doped samples, with systematically varying intensity and FWHM values. This new emission band can be attributed to electronic transitions involving the Cr3⁺-induced defect complexes ( C r B a , V B a , and O i ) [58].
A notable feature is the progressive decrease in overall PL intensity with increasing Cr3⁺ concentration, particularly pronounced for 3%- and 5%-doped samples, where the integrated intensity decreases to approximately 45% and 25% of the pristine value, respectively. This reduction correlates with the enhanced tetragonal distortion and increased defect concentration, suggesting that a higher Cr3⁺ content promotes non-radiative recombination pathways [59]. The relative contributions of emission components also evolve systematically, with the second peak (392–396 nm) maintaining a dominant contribution (~68–72%) in doped samples while the intensity of the third peak decreases significantly. Peak positions show subtle but consistent shifts (∆λ ≤ 3 nm) with increasing Cr3⁺ content, indicating modification of the local electronic environment around emission centers.
The evolution of PL characteristics provides valuable insights into the defect structure and electronic properties of Cr3⁺-doped BaTiO3. The appearance of the fourth emission component and the modification of existing emission intensities directly reflect the formation and interaction of defect complexes arising from A-site Cr3⁺ substitution [60]. These spectroscopic signatures, combined with the observed trends in peak parameters and overall intensity, strongly support our proposed model of defect-mediated accommodation of Cr3⁺ at the Ba2⁺ site. The high quality of peak deconvolution (R2 > 0.999) and the systematic nature of spectral changes across the doping series provide robust evidence for the proposed electronic structure modifications.

3.7. Transmission Electron Microscopy

The transmission electron microscopy (TEM) analysis elucidates the morphological and structural evolution of Cr3⁺-doped BaTiO3 nanoparticles as a function of dopant concentration (Figure 7). Quantitative size distribution analysis, based on measurements of over 100 particles per sample with uniform 10 nm interval widths (Figure S3a–f), reveals systematic changes in particle size statistics. The pristine BaTiO3 exhibits a broad size distribution with mean diameter 74.52 ± 74.04 nm and median 48.59 nm, indicating significant size polydispersity. Upon Cr3⁺ incorporation, a dramatic reduction in both average size and distribution width is observed, with the 0.3% Cr3⁺-doped sample showing a narrower distribution with mean 33.46 ± 28.82 nm and median 32.954 nm, suggesting more uniform nucleation and growth dynamics. Further doping reveals systematic evolution in particle statistics, with 0.5% Cr3⁺ showing a mean of 32.84 ± 19.29 nm (median 27.98 nm), 1.0% Cr3⁺ with a mean of 33.61 ± 17.44 nm (median 26.77 nm), 3.0% Cr3⁺ displaying a mean of 40.02 ± 25.30 nm (median 32.94 nm), and 5.0% Cr3⁺ exhibiting means of 42.71 ± 24.18 nm (median 35.89 nm). The size distributions follow a lognormal pattern, with decreasing skewness at higher doping concentrations, suggesting more controlled particle growth [15].
High-resolution TEM (HRTEM) analysis reveals well-resolved lattice fringes, enabling direct measurement of interplanar spacings. The pristine BaTiO3 (Figure 7a) shows clear lattice fringes with a d-spacing of 0.289 nm, corresponding to the (111) planes. Upon Cr3⁺ doping, systematic variations in the lattice parameters are observed. The 0.3% Cr3⁺-doped sample (Figure 7b) exhibits lattice fringes with d = 0.393 nm indexed to (100) planes, while the 0.5% Cr3⁺-doped sample (Figure 7c) shows fringes with d = 0.279 nm corresponding to (110) planes. The 1%, 3%, and 5% Cr3⁺-doped samples (Figure 7d–f) display interplanar spacings of 0.394 nm, all corresponding to (100) planes.
Energy-dispersive X-ray spectroscopy (EDX) analysis (Table 2) quantitatively confirms the successful and controlled incorporation of Cr3⁺, with atomic concentrations incrementing from 0.46 atom% for 0.3% nominal Cr3⁺-doping to 1.64 atom% for 5% nominal Cr3⁺-doping. The remarkable reduction in particle size upon initial Cr3⁺ doping (0.3–1.0%) suggests that even small amounts of dopant significantly alter the nucleation dynamics. The subsequent gradual increase in particle size at higher doping levels (3–5% Cr3⁺) indicates a transition in the growth mechanism, possibly due to the increased concentration of defect complexes affecting crystal growth kinetics [61].
The convergence of mean and median values with increasing Cr3⁺ content indicates enhanced size uniformity, particularly evident in the 3% and 5% doped samples, suggesting a more controlled growth process mediated by the presence of dopant-induced defect complexes [15].

3.8. DPPH Radical Antioxidant Activity

The antioxidant activity of pure and Cr3+-doped BaTiO3 nanoparticles was evaluated using the DPPH assay, which measures the ability of the nanoparticles to scavenge stable free radicals. Figure 8 shows that the scavenging percentages obtained from the assay provide valuable insights into the radical scavenging potential of the nanoparticles at different concentrations (125, 250, 500, and 1000 μg/mL). Pure BaTiO3 nanoparticles exhibited relatively stable scavenging percentages across the tested concentrations, ranging from 40.70% to 43.33%, indicating their inherent antioxidant properties. On the other hand, Cr3+-doped BaTiO3 nanoparticles displayed a more pronounced variation in scavenging percentages, particularly at higher concentrations. At lower concentrations (125 and 250 μg/mL), the Cr3+-doped samples showed scavenging percentages comparable to or slightly lower than pure BaTiO3. However, as the concentration increased to 500 and 1000 μg/mL, the scavenging percentages of Cr3+-doped samples tended to decrease, especially for higher doping concentrations (3% and 5% Cr3+). This trend suggests that the incorporation of Cr3+ into the BaTiO3 lattice influences the antioxidant activity of the nanoparticles, likely due to structural modifications, electronic effects, and changes in surface chemistry. The structural transition from cubic to tetragonal phase induced by Cr3+ doping may affect the surface properties and reactivity of the nanoparticles [49], impacting their ability to interact with and neutralize DPPH radicals [62]. Additionally, the alteration of the electronic structure and band gap of the material, as well as the presence of Cr3+ ions on the surface, may influence the electron transfer processes involved in the scavenging of DPPH radicals [63], leading to a decrease in antioxidant activity at higher doping concentrations.
A comprehensive analysis of the antioxidant activity of Cr3+-doped BaTiO3 nanoparticles was conducted using a two-way analysis of variance (ANOVA) (Table 3). The results revealed highly significant main effects for both the sample type (F(5, 48) = 47.45, p < 0.001) and DPPH concentration (F(3, 48) = 58.70, p < 0.001). Furthermore, a strong interaction effect between sample type and DPPH concentration was observed (F(15, 48) = 23.99, p < 0.001). These findings provide robust evidence that the level of Cr3+ doping in BaTiO3 nanoparticles significantly influences their antioxidant activity, and this effect is markedly dependent on the DPPH concentration used in the assay. The substantial interaction effect underscores a complex relationship between Cr3+ doping and DPPH concentration, highlighting the critical importance of considering both factors when evaluating the antioxidant properties of these nanoparticles. The use of triplicate measurements enhances the reliability of these results.

3.9. Antibacterial Activity

The antibacterial activity of pure and Cr3+-doped BaTiO3 nanoparticles against E. coli, evaluated using the Kirby–Bauer disk diffusion method (Figure S4), reveals a complex relationship between the Cr3+ doping concentration and antibacterial efficacy (Table 4). Deionized water, used as a negative control, showed no inhibition zone, confirming the specificity of the observed antibacterial effects to the nanoparticle treatments. Pure BaTiO3 nanoparticles exhibited stable inhibition zone diameters (8.988–9.886 mm) across the tested concentrations (0.5–4 mg/mL), indicating inherent antibacterial properties attributed to reactive oxygen species (ROS) generation and bacterial membrane disruption [64]. Cr3+ doping significantly influenced this activity, with a non-linear trend observed. At low Cr3+ concentrations (0.3% and 0.5%), antibacterial activity was enhanced, particularly at higher nanoparticle concentrations (1.5–4 mg/mL), with the 0.5% Cr3+-doped sample showing peak performance (10.569 mm at 1.5 mg/mL). This enhancement likely stems from optimal structural distortion, as evidenced by XRD results showing the cubic-to-tetragonal transition, combined with the introduction of non-radiative recombination centers observed in the PL spectra, which may facilitate enhanced ROS generation and bacterial interaction [65]. However, at higher Cr3+ concentrations (1%, 3%, and 5%), the antibacterial efficacy decreased, with 5% Cr3+-doping exhibiting the lowest activity (6.892–8.992 mm). This decline can be attributed to several factors: excessive structural distortion, as indicated by XRD and Raman spectroscopy, potentially affecting surface reactivity and bacterial membrane interaction; electronic structure modifications, revealed by PL and UV–vis spectroscopy, possibly leading to less efficient ROS generation; and surface property changes, evidenced by the broadening of PL emission peaks and TEM observations of increased surface roughness, potentially reducing active sites for bacterial interaction. The incorporation of Cr3+ ions into the BaTiO3 lattice alters the electronic structure and band gap of the material [38,66], influencing the generation and transfer of ROS, which play a crucial role in the antibacterial mechanism of metal oxide nanoparticles. At higher Cr3+ doping concentrations, these electronic modifications may become less favorable for efficient ROS generation and transfer, thereby reducing antibacterial efficacy. Additionally, the presence of Cr3+ ions on the nanoparticle surface may modify surface chemistry and the availability of reactive sites for bacterial cell interaction, potentially leading to a saturation of surface sites at higher doping levels and reducing overall antibacterial efficiency. The observed trends align with the complex interplay between structural transitions, electronic structure modifications, and surface property changes induced by Cr3+ doping. The optimal 0.5% Cr3+ concentration likely represents a balance point where these factors synergistically enhance antibacterial activity, while higher concentrations lead to detrimental effects.

4. Conclusions

This comprehensive study on Cr3⁺-doped BaTiO3 nanoparticles demonstrates that controlled Cr3⁺ doping induces significant modifications across multiple material properties. The sol–gel synthesis yields well-crystallized nanoparticles with a systematic size evolution, showing an initial decrease from 74.52 nm to 33.46 nm at 0.3% Cr3⁺ followed by a gradual increase to 42.71 nm at 5% Cr3⁺. Structural characterization reveals a cubic-to-tetragonal phase transition upon doping, accompanied by systematic changes in lattice parameters and the formation of defect complexes. The optical properties show remarkable tunability, with bandgap reduction from 3.168 eV to 2.604 eV and systematic modifications in photoluminescence characteristics, including the emergence of new emission components and controlled intensity variations. Vibrational spectroscopy confirms the structural transitions through systematic changes in Raman and FTIR modes. High-resolution TEM analysis reveals systematic variations in lattice spacings and morphology, while EDX confirms successful dopant incorporation up to 1.64 atom%. The bioactive properties exhibit concentration-dependent behavior, with optimal antibacterial activity at 0.5% Cr3⁺ doping (10.569 mm inhibition zone) and sustained antioxidant activity in pure and Cr3+-doped BaTiO3. These findings establish clear structure–property relationships in Cr3⁺-doped BaTiO3 nanoparticles and demonstrate their potential for applications ranging from electronics to biomedicine. Future research directions should include investigations into long-term stability, exploration of co-doping strategies, and in-depth mechanistic studies to further optimize these multifunctional nanomaterials for next-generation applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst14110998/s1, Figure S1: COD X-ray diffraction standard; Figure S2: Rietveld refinement plots; Figure S3: Influence of dopant concentration on BaTiO3 lattice constant and the tetragonality; Figure S4: Deconvolution peaks; Figure S5: TEM particle size distribution and micrographs; Figure S6: Inhition zone; Table S1: Deconvolution parameters.

Author Contributions

Conceptualization, E.M.F. and J.A.C.G.; methodology J.C.H.H., E.J.S.S., E.M.F. and J.A.C.G.; software, E.J.S.S.; validation, C.A.P.B. and H.E.N.M.; formal analysis, J.C.H.H. and H.E.N.M.; investigation, J.C.H.H., R.M.T.C., B.S.V.B. and G.d.L.F.Q.; resources, F.G.G. and A.B.Q.C.; data curation, J.C.H.H., G.d.L.F.Q. and C.A.P.B.; writing—original draft preparation, J.C.H.H., B.S.V.B. and E.J.S.S.; writing—review and editing, E.M.F., B.S.V.B., A.B.Q.C. and E.J.S.S.; visualization, J.C.H.H., G.d.L.F.Q. and E.J.S.S.; supervision, C.A.P.B. and E.M.F.; project administration, E.M.F. and J.A.C.G.; funding acquisition, E.M.F., J.A.C.G. and B.S.V.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Universidad Nacional Jorge Basadre Grohmann through “Fondos del canon sobrecanon y regalias mineras”, with the projects “Estudio de materiales ferroeléctricos (BiFeO3 y Bi2FeCrO6)” and “Determinación de huellas digitales ópticas, de materiales sólidos, líquidos y orgánicos mediante espectroscopia visible e infrarroja” approved with the resolutions N° 5854-2019-UN/JBG and N° 4516-2018-UN/JBG, respectively. The project “Caracterización microestructural de un antibiótico boratado, obtenido de boro residual, y su aplicación en el diseño de recubrimientos fotocatalíticos urbanos antibacterianos” was approved through resolution R.C.O N°. 352-2022-UNAM.

Data Availability Statement

The author confirms that all obtained during this study are included in this published article and its supplementary material.

Acknowledgments

We are grateful to Universidad Nacional Jorge Basadre Grohman and the projects, “Estudio de materiales ferroeléctricos (BiFeO3 y Bi2FeCrO6) y su aplicación en celdas solares”, “Determinación de huellas digitales ópticas, de materiales sólidos, líquidos y orgánicos mediante espectroscopia visible e infrarroja”, and “Generación fotocatalítica y foto-electrocatalítica de hidrógeno en la región Tacna empleando nanopartículas de NiTiO3 puras y dopadas”, approved by rectoral resolutions N° 4516-2018-UN/JBG, N° 5854-2019-UN/JBG, and N° 9155-2021-UN/JBG, respectively, for their support. The author Bertha Silvana Vera Barrios expresses her gratitude to the National University of Moquegua and to her project “Caracterización microestructural de un antibiótico boratado, obtenido de boro residual, y su aplicación en el diseño de recubrimientos fotocatalíticos urbanos antibacterianos”, which was approved through resolution R.C.O N°. 352-2022-UNAM. The author Elisban Juani Sacari Sacari gratefully acknowledges the financial support provided by CONCYTEC through the PROCIENCIA program under the “Becas en programas de doctorado en alianzas interinstitucionales” competition, according to contracts N°PE501088673-2024-PROCIENCIA-BM and N°PE501084296-2023-PROCIENCIA-BM for undertaking a Doctoral program in Physics at the Universidad Nacional de Ingeniería, Peru.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Jain, A.; Kumar, A.; Gupta, N.; Kumar, K. Advancements and challenges in BaTiO3-Based materials for enhanced energy storage. Mater. Today Proc. 2024, in press. [Google Scholar] [CrossRef]
  2. Masekela, D.; Hintsho-Mbita, N.C.; Sam, S.; Yusuf, T.L.; Mabuba, N. Application of BaTiO3-based catalysts for piezocatalytic, photocatalytic and piezo-photocatalytic degradation of organic pollutants and bacterial disinfection in wastewater: A comprehensive review. Arab. J. Chem. 2023, 16, 104473. [Google Scholar] [CrossRef]
  3. Sood, A.; Desseigne, M.; Dev, A.; Maurizi, L.; Kumar, A.; Millot, N.; Han, S.S. A Comprehensive Review on Barium Titanate Nanoparticles as a Persuasive Piezoelectric Material for Biomedical Applications: Prospects and Challenges. Small 2023, 19, e2206401. [Google Scholar] [CrossRef] [PubMed]
  4. Lang, X.Y.; Jiang, Q. Size and interface effects on Curie temperature of perovskite ferroelectric nanosolids. J. Nanopart. Res. 2007, 9, 595–603. [Google Scholar] [CrossRef]
  5. Jiang, Q.; Cui, X.F.; Zhao, M. Size effects on Curie temperature of ferroelectric particles. Appl. Phys. A Mater. Sci. Process. 2004, 78, 703–704. [Google Scholar] [CrossRef]
  6. Wada, S.; Hoshina, T.; Yasuno, H.; Ohishi, M.; Kakemoto, H.; Tsurumi, T.; Yashima, M. Size Effect of Dielectric Properties for Barium Titanate Particles and Its Model. Key Eng. Mater. 2006, 301, 27–30. [Google Scholar] [CrossRef]
  7. Shen, Z.-Y.; Li, J.-F. Enhancement of piezoelectric constant d33 in BaTiO3 ceramics due to nano-domain structure. J. Ceram. Soc. Jpn. 2010, 118, 940–943. [Google Scholar] [CrossRef]
  8. Lu, W.; Quilitz, M.; Schmidt, H. Nanoscaled BaTiO3 powders with a large surface area synthesized by precipitation from aqueous solutions: Preparation, characterization and sintering. J. Eur. Ceram. Soc. 2007, 27, 3149–3159. [Google Scholar] [CrossRef]
  9. Ashiri, R. On the solid-state formation of BaTiO3 nanocrystals from mechanically activated BaCO3 and TiO2 powders: Innovative mechanochemical processing, the mechanism involved, and phase and nanostructure evolutions. RSC Adv. 2016, 6, 17138–17150. [Google Scholar] [CrossRef]
  10. Suherman, B.; Nurosyid, F.; Khairuddin; Sandi, D.K.; Irian, Y. Impacts of low sintering temperature on microstructure, atomic bonds, and dielectric constant of barium titanate (BaTiO3) prepared by co-precipitation technique. J. Phys. Conf. Ser. 2022, 2190, 12006. [Google Scholar] [CrossRef]
  11. Hayashi, H.; Hakuta, Y. Hydrothermal Synthesis of Metal Oxide Nanoparticles in Supercritical Water. Materials 2010, 3, 3794–3817. [Google Scholar] [CrossRef] [PubMed]
  12. Khort, A.A.; Podbolotov, K.B. Preparation of BaTiO3 nanopowders by the solution combustion method. Ceram. Int. 2016, 42, 15343–15348. [Google Scholar] [CrossRef]
  13. Choi, G.J.; Kim, H.S.; Cho, Y.S. BaTiO3 particles prepared by microwave-assisted hydrothermal reaction using titanium acylate precursors. Mater. Lett. 1999, 41, 122–127. [Google Scholar] [CrossRef]
  14. Kavian, R.; Saidi, A. Sol–gel derived BaTiO3 nanopowders. J. Alloys Compd. 2009, 468, 528–532. [Google Scholar] [CrossRef]
  15. Navas, D.; Fuentes, S.; Castro-Alvarez, A.; Chavez-Angel, E. Review on Sol-Gel Synthesis of Perovskite and Oxide Nanomaterials. Gels 2021, 7, 275. [Google Scholar] [CrossRef]
  16. Borlaf, M.; Moreno, R. Colloidal sol-gel: A powerful low-temperature aqueous synthesis route of nanosized powders and suspensions. Open Ceram. 2021, 8, 100200. [Google Scholar] [CrossRef]
  17. Rahman, M.A. Understanding of doping sites and versatile applications of heteroatom modified BaTiO3 ceramic. J. Asian Ceram. Soc. 2023, 11, 215–224. [Google Scholar] [CrossRef]
  18. Issam, D.; Achehboune, M.; Boukhoubza, I.; Hatel, R.; El Adnani, Z.; Rezzouk, A. Investigation of the crystal structure, electronic and optical properties of Cr-doped BaTiO3 on the Ti site using first principles calculations. J. Phys. Chem. Solids 2023, 175, 111209. [Google Scholar] [CrossRef]
  19. Derkaoui, I.; Achehboune, M.; Boukhoubza, I.; El Adnani, Z.; Rezzouk, A. Improved first-principles electronic band structure for cubic (Pm 3 m) and tetragonal (P4mm, P4/mmm) phases of BaTiO3 using the Hubbard U correction. Comput. Mater. Sci. 2023, 217, 111913. [Google Scholar] [CrossRef]
  20. Buscaglia, V.; Buscaglia, M.T.; Canu, G. BaTiO3-Based Ceramics: Fundamentals, Properties and Applications. Encyclopedia of Materials: Technical Ceramics and Glasses; Elsevier: Amsterdam, The Netherlands, 2021; pp. 311–344. ISBN 9780128222331. [Google Scholar]
  21. Amaechi, I.C.; Kolhatkar, G.; Youssef, A.H.; Rawach, D.; Sun, S.; Ruediger, A. B-site modified photoferroic Cr3+-doped barium titanate nanoparticles: Microwave-assisted hydrothermal synthesis, photocatalytic and electrochemical properties. RSC Adv. 2019, 9, 20806–20817. [Google Scholar] [CrossRef]
  22. Ramakanth, S.; James Raju, K.C. Band gap narrowing in BaTiO3 nanoparticles facilitated by multiple mechanisms. J. Appl. Phys. 2014, 115, 173507. [Google Scholar] [CrossRef]
  23. Srilakshmi, C.; Saraf, R.; Prashanth, V.; Rao, G.M.; Shivakumara, C. Structure and Catalytic Activity of Cr-Doped BaTiO3 Nanocatalysts Synthesized by Conventional Oxalate and Microwave Assisted Hydrothermal Methods. Inorg. Chem. 2016, 55, 4795–4805. [Google Scholar] [CrossRef] [PubMed]
  24. Tewatia, K.; Sharma, A.; Sharma, M.; Kumar, A. Factors affecting morphological and electrical properties of Barium Titanate: A brief review. Mater. Today Proc. 2021, 44, 4548–4556. [Google Scholar] [CrossRef]
  25. Ray, S.K.; Cho, J.; Hur, J. A critical review on strategies for improving efficiency of BaTiO3-based photocatalysts for wastewater treatment. J. Environ. Manag. 2021, 290, 112679. [Google Scholar] [CrossRef] [PubMed]
  26. Kumar, A.; Gori, Y.; Kumar, A.; Meena, C.S.; Dutt, N. (Eds.) Advanced Materials for Biomedical Applications, 1st ed.; Taylor and Francis: Boca Raton, FL, USA, 2023; ISBN 9781032356068. [Google Scholar]
  27. Lay, R.; Deijs, G.S.; Malmström, J. The intrinsic piezoelectric properties of materials—A review with a focus on biological materials. RSC Adv. 2021, 11, 30657–30673. [Google Scholar] [CrossRef]
  28. Tiburcio, J.; Sacari, E.; Chacaltana, J.; Medina, J.; Gamarra, F.; Polo, C.; Mamani, E.; Quispe, A. Influence of Cr Doping on Structural, Optical, and Photovoltaic Properties of BiFeO3 Synthesized by Sol-Gel Method. Energies 2023, 16, 786. [Google Scholar] [CrossRef]
  29. Anjos, P.d.; Pereira, E.C.; Gobato, Y.G. Study of the structure and optical properties of rare-earth-doped aluminate particles prepared by an amorphous citrate sol–gel process. J. Alloys Compd. 2005, 391, 277–283. [Google Scholar] [CrossRef]
  30. Kao, C.-F.; Yang, W.-D. Preparation of barium strontium titanate powder from citrate precursor. Appl. Organometal. Chem. 1999, 13, 383–397. [Google Scholar] [CrossRef]
  31. Pérez-Maqueda, L.A.; Diánez, M.J.; Gotor, F.J.; Sayagués, M.J.; Real, C.; Criado, J.M. Synthesis of needle-like BaTiO3 particles from the thermal decomposition of a citrate precursor under sample controlled reaction temperature conditions. J. Mater. Chem. 2003, 13, 2234–2241. [Google Scholar] [CrossRef]
  32. Benyoussef, M.; Mura, T.; Saitzek, S.; Azrour, F.; Blach, J.-F.; Lahmar, A.; Gagou, Y.; El Marssi, M.; Sayede, A.; Jouiad, M. Nanostructured BaTi1−xSnxO3 ferroelectric materials for electrocaloric applications and energy performance. Curr. Appl. Phys. 2022, 38, 59–66. [Google Scholar] [CrossRef]
  33. Sugawara, K.; Sakusabe, H.; Nishino, T.; Sugawara, T.; Ogiwara, K.; Dranoff, J.S. Thermal decomposition of barium titanate precursor prepared by a wet chemical method. AIChE J. 1997, 43, 2837–2843. [Google Scholar] [CrossRef]
  34. Elmehdi, H.M.; Ramachandran, K.; Chidambaram, S.; Mani, G.T.; Pandiaraj, S.; Alqarni, S.A.; Daoudi, K.; Gaidi, M. Diode characteristics, piezo-photocatalytic antibiotic degradation and hydrogen production of Ce3+ doped ZnO nanostructures. Chemosphere 2024, 350, 141015. [Google Scholar] [CrossRef] [PubMed]
  35. Qiao, L.; Bi, X. Microstructure and grain size dependence of ferroelectric properties of BaTiO3 thin films on LaNiO3 buffered Si. J. Eur. Ceram. Soc. 2009, 29, 1995–2001. [Google Scholar] [CrossRef]
  36. Gražulis, S.; Daškevič, A.; Merkys, A.; Chateigner, D.; Lutterotti, L.; Quirós, M.; Serebryanaya, N.R.; Moeck, P.; Downs, R.T.; Le Bail, A. Crystallography Open Database (COD): An open-access collection of crystal structures and platform for world-wide collaboration. Nucleic Acids Res. 2012, 40, D420-7. [Google Scholar] [CrossRef]
  37. Silva, R.S.; Cunha, F.; Barrozo, P. Raman spectroscopy of the Al-doping induced structural phase transition in LaCrO3 perovskite. Solid State Commun. 2021, 333, 114346. [Google Scholar] [CrossRef]
  38. Buscaglia, M.T.; Buscaglia, V.; Viviani, M.; Nanni, P.; Hanuskova, M. Influence of foreign ions on the crystal structure of BaTiO3. J. Eur. Ceram. Soc. 2000, 20, 1997–2007. [Google Scholar] [CrossRef]
  39. Khedhri, M.H.; Abdelmoula, N.; Khemakhem, H.; Douali, R.; Dubois, F. Structural, spectroscopic and dielectric properties of Ca-doped BaTiO3. Appl. Phys. A Mater. Sci. Process. 2019, 125, 193. [Google Scholar] [CrossRef]
  40. El Ghandouri, A.; Sayouri, S.; Lamcharfi, T.; Elbasset, A. Structural, microstructural and dielectric properties of Ba1−xLax Ti(1−x/4)O3 prepared by sol gel method. J. Adv. Dielect. 2019, 9, 1950026. [Google Scholar] [CrossRef]
  41. Da Lu, Y.; Han, D.D.; Liu, Q.L.; Wang, Y.D.; Sun, X.Y. Structure and Dielectric Properties of Ce and Ca Co-Doped BaTiO3 Ceramics. Key Eng. Mater. 2016, 680, 184–188. [Google Scholar] [CrossRef]
  42. Ren, P.; Wang, Q.; Wang, X.; Wang, L.; Wang, J.; Fan, H.; Zhao, G. Effects of doping sites on electrical properties of yttrium doped BaTiO3. Mater. Lett. 2016, 174, 197–200. [Google Scholar] [CrossRef]
  43. Rached, A.; Wederni, M.A.; Belkahla, A.; Dhahri, J.; Khirouni, K.; Alaya, S.; Martín-Palma, R.J. Effect of doping in the physico-chemical properties of BaTiO3 ceramics. Phys. B Condens. Matter 2020, 596, 412343. [Google Scholar] [CrossRef]
  44. Yen, F.-S.; Hsiang, H.-I.; Chang, Y.-H. Cubic to Tetragonal Phase Transformation of Ultrafine BaTiO3 Crystallites at Room Temperature. Jpn. J. Appl. Phys. 1995, 34, 6149. [Google Scholar] [CrossRef]
  45. Hayashi, H.; Nakamura, T.; Ebina, T. In-situ Raman spectroscopy of BaTiO3 particles for tetragonal–cubic transformation. J. Phys. Chem. Solids 2013, 74, 957–962. [Google Scholar] [CrossRef]
  46. He, Q.; Tang, X.; Zhang, J.; Wu, M. Raman study for BaTiO3 system doped with various concentrations and treated at different temperatures. Nanostructured Mater. 1999, 11, 287–293. [Google Scholar] [CrossRef]
  47. Yu, J.; Meng, X.J.; Sun, J.L.; Wang, G.S.; Chu, J.H. Phase transformation and Raman spectra in BaTiO3 nanocrystals. MRS Proc. 2002, 718, D10-7. [Google Scholar] [CrossRef]
  48. Shu, C.; Reed, D.; Button, T.W. A phase diagram of Ba1−xCaxTiO3 (x = 0–0.30) piezoceramics by Raman spectroscopy. J. Am. Ceram. Soc. 2018, 101, 2589–2593. [Google Scholar] [CrossRef]
  49. Kappadan, S.; Gebreab, T.W.; Thomas, S.; Kalarikkal, N. Tetragonal BaTiO3 nanoparticles: An efficient photocatalyst for the degradation of organic pollutants. Mater. Sci. Semicond. Process. 2016, 51, 42–47. [Google Scholar] [CrossRef]
  50. Del López, M.C.B.; Fourlaris, G.; Rand, B.; Riley, F.L. Characterization of Barium Titanate Powders: Barium Carbonate Identification. J. Am. Ceram. Soc. 1999, 82, 1777–1786. [Google Scholar] [CrossRef]
  51. Ravanamma, R.; Muralidhara Reddy, K.; Venkata Krishnaiah, K.; Ravi, N. Structure and morphology of yttrium doped barium titanate ceramics for multi-layer capacitor applications. Mater. Today Proc. 2021, 46, 259–262. [Google Scholar] [CrossRef]
  52. Lin, F.; Jiang, D.; Ma, X.; Shi, W. Influence of doping concentration on room-temperature ferromagnetism for Fe-doped BaTiO3 ceramics. J. Magn. Magn. Mater. 2008, 320, 691–694. [Google Scholar] [CrossRef]
  53. Jin, X.; Sun, D.; Zhang, M.; Zhu, Y.; Qian, J. Investigation on FTIR spectra of barium calcium titanate ceramics. J. Electroceram 2009, 22, 285–290. [Google Scholar] [CrossRef]
  54. Yang, F.; Yang, L.; Ai, C.; Xie, P.; Lin, S.; Wang, C.-Z.; Lu, X. Tailoring Bandgap of Perovskite BaTiO₃ by Transition Metals Co-Doping for Visible-Light Photoelectrical Applications: A First-Principles Study. Nanomaterials 2018, 8, 455. [Google Scholar] [CrossRef] [PubMed]
  55. Tian, M.; Han, A.; Ma, S.; Zhu, X.; Ye, M.; Chen, X. Preparation of Cr-doped BaTiO3 near infrared reflection pigment powder and its anti-aging performance for acrylonitrile-styrene-acrylate. Powder Technol. 2021, 378, 182–190. [Google Scholar] [CrossRef]
  56. Eden, S.; Kapphan, S.; Hesse, H.; Trepakov, V.; Vikhnin, V.; Gregora, I.; Jastrabik, L.; Seglins, J. Observations of the absorption, infra-red emission, and excitation spectra of Cr in. J. Phys. Condens. Matter 1998, 10, 10775–10786. [Google Scholar] [CrossRef]
  57. Orhan, E.; Pontes, F.M.; Pinheiro, C.D.; Longo, E.; Pizani, P.S.; Varela, J.A.; Leite, E.R.; Boschi, T.M.; Beltrán, A.; Andrés, J. Theoretical and experimental study of the relation between photoluminescence and structural disorder in barium and strontium titanate thin films. J. Eur. Ceram. Soc. 2005, 25, 2337–2340. [Google Scholar] [CrossRef]
  58. Wang, Y.; Zhou, Q.; Zhang, Q.; Ren, Y.; Cui, K.; Cheng, C.; Wu, K. Effects of La-N Co-Doping of BaTiO3 on Its Electron-Optical Properties for Photocatalysis: A DFT Study. Molecules 2024, 29, 2250. [Google Scholar] [CrossRef]
  59. Rahman, M.A.; Hasan, Z.; Islam, J.; Das, D.K.; Chowdhury, F.I.; Khandaker, M.U.; Zabed, H.M.; Bradley, D.A.; Osman, H.; Ullah, M.H. Tailoring the Properties of Bulk BaTiO3 Based Perovskites by Heteroatom-Doping towards Multifunctional Applications: A Review. ECS J. Solid State Sci. Technol. 2023, 12, 103015. [Google Scholar] [CrossRef]
  60. Perovskite Metal Oxides; Elsevier: Amsterdam, The Netherlands, 2023; ISBN 9780323995290.
  61. Hsiang, H.-I.; Yen, F.-S.; Chang, Y.-H. Effects of doping with La and Mn on the crystallite growth and phase transition of BaTiO3 powders. J. Mater. Sci. 1996, 31, 2417–2424. [Google Scholar] [CrossRef]
  62. Li, L.; Guo, R.; Gao, J.; Liu, J.; Zhao, Z.; Sheng, X.; Fan, J.; Cui, F. Insight into mechanochemical destruction of PFOA by BaTiO3: An electron-dominated reduction process. J. Hazard. Mater. 2023, 450, 131028. [Google Scholar] [CrossRef]
  63. Sanaullah, I.; Khan, H.N.; Sajjad, A.; Khan, S.; Sabri, A.N.; Naseem, S.; Riaz, S. Improved osteointegration response using high strength perovskite BaTiO3 coatings prepared by chemical bath deposition. J. Mech. Behav. Biomed. Mater. 2023, 138, 105635. [Google Scholar] [CrossRef]
  64. Boschetto, F.; Doan, H.N.; Phong Vo, P.; Zanocco, M.; Yamamoto, K.; Zhu, W.; Adachi, T.; Kinashi, K.; Marin, E.; Pezzotti, G. Bacteriostatic Behavior of PLA-BaTiO3 Composite Fibers Synthesized by Centrifugal Spinning and Subjected to Aging Test. Molecules 2021, 26, 2918. [Google Scholar] [CrossRef] [PubMed]
  65. Kumar, A.; Kumar, S.; Prajapat, M.; Prakash, C. Magnetodielectric properties of Cr3+ ions doped BaTiO3 multiferroic ceramic. In Proceedings of the Nanoforum 2014, Rome, Italy, 22–25 September 2014; AIP Publishing LLC.: Melville, NY, USA, 2015; p. 140008. [Google Scholar]
  66. Hossain, K.M.; Ahmad, S.; Mitro, S.K. Physical properties of chromium-doped barium titanate: Effects of chromium incorporation. Phys. B Condens. Matter 2022, 626, 413494. [Google Scholar] [CrossRef]
Figure 1. (a) Thermogravimetric analysis and (b) differential scanning calorimetry of pure and Cr3+-doped BaTiO3.
Figure 1. (a) Thermogravimetric analysis and (b) differential scanning calorimetry of pure and Cr3+-doped BaTiO3.
Crystals 14 00998 g001
Figure 2. (a) X-ray diffraction patterns of pure and doped BaTiO3 and (b) amplification of XRD peaks within the 44–46.5° range.
Figure 2. (a) X-ray diffraction patterns of pure and doped BaTiO3 and (b) amplification of XRD peaks within the 44–46.5° range.
Crystals 14 00998 g002
Figure 3. Raman spectra of (a) Pristine BaTiO3 and (b) Cr3+-doped BaTiO3.
Figure 3. Raman spectra of (a) Pristine BaTiO3 and (b) Cr3+-doped BaTiO3.
Crystals 14 00998 g003
Figure 4. FTIR spectra of pure and Cr3+-doped BaTiO3.
Figure 4. FTIR spectra of pure and Cr3+-doped BaTiO3.
Crystals 14 00998 g004
Figure 5. (a) UV–visible diffuse reflectance spectrum. (b) Kubelka–Munk plot for bandgap calculation.
Figure 5. (a) UV–visible diffuse reflectance spectrum. (b) Kubelka–Munk plot for bandgap calculation.
Crystals 14 00998 g005
Figure 6. Photoluminescence spectrums of pure and Cr3+-doped BaTiO3.
Figure 6. Photoluminescence spectrums of pure and Cr3+-doped BaTiO3.
Crystals 14 00998 g006
Figure 7. Transmission electron microscopy microphotography of (a) BaTiO3, (b) BaTiO3-0.3%Cr3+, (c) BaTiO3-0.5%Cr3+, (d) BaTiO3-1%Cr3+, (e) BaTiO3-3%Cr3+, and (f) BaTiO3-5%Cr3+.
Figure 7. Transmission electron microscopy microphotography of (a) BaTiO3, (b) BaTiO3-0.3%Cr3+, (c) BaTiO3-0.5%Cr3+, (d) BaTiO3-1%Cr3+, (e) BaTiO3-3%Cr3+, and (f) BaTiO3-5%Cr3+.
Crystals 14 00998 g007
Figure 8. Antioxidant activity of pure and Cr3+-doped BaTiO3.
Figure 8. Antioxidant activity of pure and Cr3+-doped BaTiO3.
Crystals 14 00998 g008
Table 1. Structural parameters of chromium-doped BaTiO3 nanoparticles.
Table 1. Structural parameters of chromium-doped BaTiO3 nanoparticles.
Structural ParameterSample
BaTiO3BaTiO3-0.3%Cr3+BaTiO3-0.5%Cr3+BaTiO3-1%Cr3+BaTiO3-3%Cr3+BaTiO3-5%Cr3+
Crystal systemCubicTetragonalTetragonalTetragonalTetragonalTetragonal
Space groupPm-3mP4mmP4mmP4mmP4mmP4mm
A = b (Å)4.007854.006984.005544.005994.004764.00379
c (Å)4.007854.014774.016024.015554.015584.01699
α = β = γ (°)909090909090
ρ(g/cm3)6.016.016.0166.016.01
D (nm)63.2755.2954.4954.6157.7158.27
Micro strain (%)0.0970.0630.0560.0620.0530.046
Rexp (%)2.253362.374182.3863.252212.365592.34574
Rp (%)2.823953.213064.257223.532443.243473.31153
Rwp (%)4.086844.306876.037944.507864.294434.37974
GOF1.813671.814052.530571.386091.815371.8671
Table 2. Energy dispersive X-ray spectroscopy (EDX) analysis of all samples.
Table 2. Energy dispersive X-ray spectroscopy (EDX) analysis of all samples.
ElementSample
BaTiO3BaTiO3-0.3%Cr3+BaTiO3-0.5%Cr3+BaTiO3-1%Cr3+BaTiO3-3%Cr3+BaTiO3-5%Cr3+
Ba (Atom %)18.0614.8116.3516.9117.6213.34
Ti (Atom %)18.1315.3917.6717.5718.8714.92
O (Atom %)58.5352.6651.1750.1647.0853.61
Cr (Atom %)-0.460.570.971.321.64
C (Atom%)5.2816.6814.2414.3915.1116.49
Total (%)100100100100100100
Table 3. ANOVA results.
Table 3. ANOVA results.
Source of VariationDegrees of FreedomSum of SquaresMean SquareF-Valuep-Value
Sample5362.7672.55247.45<0.001
Concentration3269.2889.7658.7<0.001
Sample × Concentration15550.2336.68223.99<0.001
Residuals4873.411.529
Table 4. Inhibition zones at various concentrations of pure and doped BaTiO3.
Table 4. Inhibition zones at various concentrations of pure and doped BaTiO3.
Zone of Inhibition (mm)
Concentration (mg/mL)0.5SD1SD1.5SD2SD4SD
BaTiO39.6020.3189.8860.3239.7070.2738.9880.3289.4070.122
BaTiO3-0.3%Cr3+9.5060.0928.1810.1198.3190.1438.8930.2939.8990.309
BaTiO3-0.5%Cr3+9.5250.3489.8150.14510.5690.1949.3090.34410.5100.219
BaTiO3-1%Cr3+8.2960.1378.2670.1318.8640.2989.7210.0798.3840.267
BaTiO3-3%Cr3+8.8700.3078.7450.1479.1230.2999.5820.1018.9400.073
BaTiO3-5%Cr3+6.8920.1277.2500.1487.8020.2388.4040.1438.9920.185
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mamani Flores, E.; Vera Barrios, B.S.; Huillca Huillca, J.C.; Chacaltana García, J.A.; Polo Bravo, C.A.; Nina Mendoza, H.E.; Quispe Cohaila, A.B.; Gamarra Gómez, F.; Tamayo Calderón, R.M.; Fora Quispe, G.d.L.; et al. Cr3+ Doping Effects on Structural, Optical, and Morphological Characteristics of BaTiO3 Nanoparticles and Their Bioactive Behavior. Crystals 2024, 14, 998. https://doi.org/10.3390/cryst14110998

AMA Style

Mamani Flores E, Vera Barrios BS, Huillca Huillca JC, Chacaltana García JA, Polo Bravo CA, Nina Mendoza HE, Quispe Cohaila AB, Gamarra Gómez F, Tamayo Calderón RM, Fora Quispe GdL, et al. Cr3+ Doping Effects on Structural, Optical, and Morphological Characteristics of BaTiO3 Nanoparticles and Their Bioactive Behavior. Crystals. 2024; 14(11):998. https://doi.org/10.3390/cryst14110998

Chicago/Turabian Style

Mamani Flores, Efracio, Bertha Silvana Vera Barrios, Julio César Huillca Huillca, Jesús Alfredo Chacaltana García, Carlos Armando Polo Bravo, Henry Edgardo Nina Mendoza, Alberto Bacilio Quispe Cohaila, Francisco Gamarra Gómez, Rocío María Tamayo Calderón, Gabriela de Lourdes Fora Quispe, and et al. 2024. "Cr3+ Doping Effects on Structural, Optical, and Morphological Characteristics of BaTiO3 Nanoparticles and Their Bioactive Behavior" Crystals 14, no. 11: 998. https://doi.org/10.3390/cryst14110998

APA Style

Mamani Flores, E., Vera Barrios, B. S., Huillca Huillca, J. C., Chacaltana García, J. A., Polo Bravo, C. A., Nina Mendoza, H. E., Quispe Cohaila, A. B., Gamarra Gómez, F., Tamayo Calderón, R. M., Fora Quispe, G. d. L., & Sacari Sacari, E. J. (2024). Cr3+ Doping Effects on Structural, Optical, and Morphological Characteristics of BaTiO3 Nanoparticles and Their Bioactive Behavior. Crystals, 14(11), 998. https://doi.org/10.3390/cryst14110998

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop