In the industrial sector, productive and effective treatment of toxic dye-based color pollutants is a key issue. Lanthanum and neodymium substituted cobalt–strontium (Co–Sr) spinel ferrite (Co0.5Sr0.5RExFe2-xO4, x = 0.00 and 0.06) catalysts were synthesized and used to degrade Congo red and rhodamine B dyes from an aqueous solution mixture in this study. For this specific purpose, RE3+ ions substituted Co–Sr spinel ferrite nanoparticles with photocatalytic degradation ability were prepared through sol–gel method. The degradation of CR and RhB in recently synthesized nanoferrites was also examined. SEM and XRD were used to characterize the prepared samples. The optical band gap values of synthesized spinel ferrites were examined with the help of Tauc plots by using UV-visible absorption. It was determined that the energy bandgap ranged from 2.91 to 2.52 eV. For Co0.5Sr0.5Fe2O4, Co0.5Sr0.5La0.06Fe1.94O4, and Co0.5Sr0.5Nd0.06Fe1.94O4 nanoferrites, the rates of CR and RhB dye degradation were 73–90% and 45–85%, respectively, at pH 5–7. The kinetics models successfully described the degradation reaction as pseudo-first-order kinetics. It was, therefore, concluded that the prepared samples can be used as effective photocatalysts in order to eliminate hazardous pollutants present in wastewater.

  • Rare-earth ion substitution spinel ferrite.

  • Ferrite for degradation of dyes.

  • Suitable band gap for photocatalytic purpose.

Despite being a minor part of the electromagnetic spectrum ranging from 100 to 400 nm, ultraviolet radiation has a significant impact on living things (Diffey 2002). Furthermore, excessive ultraviolet light can have a major impact on plant photosynthesis (Hollósy 2002; Verdaguer et al. 2017). Recently, photocatalytic degradation of dangerous chemical compounds utilizing ultraviolet light and a semiconductor material has gotten a lot of interest. This strategy is a simple way to reduce river and ecosystem pollution (Schneider et al. 2014; Di Mauro et al. 2017). So, for the purpose of water purification, degradation through photocatalysis is the most efficient solution for the elimination of a wide variety of dyes in a wide range of concentrations; hence, this application is beneficial (Mironyuk et al. 2019a, 2019b).

Spinel ferrite is considered as an important study material because of its low band gap and photocatalytic activity in the visible light area for the removal of pollutants (Niu et al. 2015a). Cobalt–Strontium (Co–Sr) spinel ferrite is a unique magnetic substance that has been used in biological, electronic, and recording technology (Bensebaa et al. 2004; Gao et al. 2009; Ahalya et al. 2014). Purity, shape, size, and attractive stability are all important factors in the photocatalytic activity and magnetic properties of ferrite nanoparticles (Haruna et al. 2020). Even though a large particle size of NPs may reduce photocatalytic activity by limiting the surface area of the catalyst, particles having smaller sizes likely to provide improved photocatalytic efficiency due to the high number of active sites (Giannakas et al. 2006). Ponraj et al. (2017) thoroughly discuss in their review article the work performed on the degradation of various dyes by using BiFeO3 NPs. Further, they examined the surface phenomena of photocatalyst BiFeO3, surface area (Wei et al. 2012; Soltani & Entezari 2013), and particle size (Gao et al. 2007; Liu et al. 2010; Gao et al. 2014; Niu et al. 2015b) and the way they impact photoactivity in their study on visible light-activated photocatalysts for the degradation of various textile dyes. The arrangement of divalent and trivalent ions in octahedral and tetrahedral locations determines the structure of spinel ferrite. Divalent compounds contain tetrahedral positions in normal spinel; however, in the inverse crystalline phase, the remaining trivalent ions occupy tetrahedral positions and the divalent ions occupy octahedral positions (Goldman 2006). Since several of the characteristics of ferromagnetic materials are reported to be influenced by the presence of impurity, significant research has been conducted that investigate the impact of cation contamination on the spinel ferrite lattice. The magnetic, electrical, and structural characteristics of ferromagnetic materials are altered when limited concentrations of these contaminants are added (Rana & Abbas 2002). Several studies have examined the effects of modifying the spinel crystalline structure of various ferromagnetic materials on their chemical characteristics as well as, as a result, magnetic permeability, their inherent characteristics of magnetization, anisotropy, electrical resistivity, Curie temperature, and magnetic resonance. Co–Sr spinel ferrite nanoparticles with lanthanum and neodymium substitutes have received a lot of attention throughout the area of nanomaterials because of their increased magnetic and electrical characteristics (Melagiriyappa & Jayanna 2009; Ishaque et al. 2010; Zhang & Wen 2010). Rare-earth ions (La3+, Nd3+) are potential replacements for enhancing the characteristics of spinel ferrites. The rare-earth element's 4f shell is insulated by 5s25p6 and is nearly completely unaffected through the prospective domain of neighboring molecules. The magnetic and electrical transport characteristics of Co–Sr spinel ferrite nanoparticles could be improved via replacing rare-earth elements and existing 4f–3d interactions (Rezlescu et al. 2000). When these are compared to undoped spinel ferrites, structural properties such as magnetization, saturation, coercivity, and retentivity as well as anisotropy, vary significantly (Kumar & Kar 2012).

Chahar et al. (2021) used the citrate precursor approach to make CoxZn1−xFe2O4 nanoferrites with varying concentrations, x = 0.00, 0.01, 0.02, 0.03, 0.04, and 0.05. It was reported that the degradation efficiency of methylene blue reached 77% for x = 0.05, compared to the minimum of 65% for x = 0.00, in 60 min during exposure to sunlight. Vosoughifar & Kimiay (2016) reported the synthesis of CuFe2−xNdxO4 photocatalyst through a sol–gel auto-combustion process and examined that methyl orange was degraded by 65% when exposed to UV light for 80 min. Naik et al. (2020) reported the fabrication of CoFe2O4 nanoparticles by sol gel route and also investigated that after 150 min of radiation, MB and EB had degraded 97 and 88%, respectively, due to low bandgap properties of prepared ferrites. Dhiman & Singhal (2019) synthesized CoRE0.02Fe1.98O4 (RE = Eu, Gd, and Dy) spinel ferrites via a simple sol–gel auto-combustion method. For the composition CoEu0.02Fe1.98O4, the maximum degradation of more than 90% was examined for both SO and RBY dyes, respectively, during 100 min of sunlight exposure. Abdo & El-Daly (2021) prepared Co0.5Cu0.5SmxFe2−xO4 (x = 0.00, 0.03, 0.06, 0.09, 0.12, and 0.15) nanoferrites by the citrate combustion approach. The nanoferrites demonstrated a degradation efficiency of 94.36% for RhB dye within 270 min at a concentration of x = 0.15, compared to Co0.5Cu0.5Fe2O4, which was 21.94%.

Co-precipitation (Maaz et al. 2009), hydrothermal method (Yan et al. 2015), and solid-state method (Deraz & Hessien 2009) are some of the methods used to make magnetic nanoparticles. The sol–gel approach, on the other hand, allows for good size control as well as a broad selection of reaction conditions for more modification (Zhang et al. 2015). The potential of sol–gel synthesis of oxide materials to produce nanoparticles exhibiting a stable pattern of crystalline solids, crystallite dimensions, and substantial surface area (Zhang et al. 2015) has piqued attention, in addition to its cheap cost and simple method.

The two dyes tested in this research, Congo red and rhodamine B, are azo dyes that are widespread in wastewater from industries and are very hazardous to human beings, causing skin irritation, nausea, blood clots, and asthma (Yu et al. 2009; Tavakoli-Azar et al. 2020). As a result, researchers have a significant task and aim to eliminate or reduce such toxins from the environment.

The goal of this study is to use the sol–gel process to synthesize pure, and rare-earth ions (La3+, Nd3+) substituted Co–Sr spinel ferrites with general formula Co0.5Sr0.5Fe2O4, Co0.5Sr0.5La0.06Fe1.94O4 and Co0.5Sr0.5Nd0.06Fe1.94O4 and to investigate its potential to degrade dyes in industrial water. Further, this study described the synthesis, characterization, and degradation efficiency of Congo red (CR) and rhodamine B (RhB) dyes.

Materials and synthesis

In a typical sol–gel method, components such as chlorides, carbonates, and nitrates are employed; however, only nitrates are utilized in this work. To synthesize an undoped Co0.5Sr0.5Fe2O4 (x = 0.00) and two doped samples with a cubic-spinel structure, Sr(NO3)2 purchased from DAEJUNG, citric acid C6H8O7, Co(NO3)2·6H2O purchased from UNI-CHEM, Fe(NO3)3·9H2O, La(NO3)3·6H2O, and N3NdO9·6H2O (99.9%) purchased from Merck-Germany and ammonia are utilized. Nitrates are dissolved in deionized water to prepare solutions of desired samples based on stoichiometric calculations. Lanthanum and neodymium are doped at a concentration of x = 0.06 to synthesize samples with the chemical compositions Co0.5Sr0.5La0.06Fe1.94O4, and Co0.5Sr0.5Nd0.06Fe1.94O4. To form a homogeneous mixture, the samples' solutions were prepared, and a magnetic stirrer was employed for constant stirring for about 10–20 min. After mixing well and preparing appropriate amounts of solution in three separate beakers, they were placed on magnetic hotplates. To keep the pH between 7 and 8, (35%) ammonia with a molarity of 51.67 mol/L was added randomly, drop by drop. The deionized water is evaporated from the solutions by heating below 80 °C and, once it evaporates, a dark brown wet gel is formed. It was then placed in an electrical thermostatic oven for 6 h approximately at 240 °C. To get loose powder, an agate mortar and pestle are employed. To achieve the final products, the fine powder is sintered in the furnace for 9 h at 950 °C. Schematic representation for the synthesis of Co0.5Sr0.5RExFe2-xO4 (x = 0.00 and 0.06) is shown in Figure 1”.

Characterization techniques

The XRD patterns of prepared spinel ferrites were accumulated over the range from 20° to 70° of 2value by XRD analysis using a Bruker D Z Phaser (PW 1830) diffractometer. The microstructures are revealed using a VEGA3 TESCAN scanning electron microscope. The optical bandgap and degradation of CR and RhB dyes for pure and rare-earth (La, Nd) doped Co–Sr spinel ferrites were determined using UV visible spectroscopy (AE-S90-2D).

Photocatalytic activity

The photocatalytic degradation of CR and RhB was carried out when exposed to sunlight. To begin, 10 ppm CR and RhB dye solutions in 100 ml of deionized water were prepared in separate beakers and stirred for 10 min with a magnetic hot plate and continuous stirring. The solution mixtures were then combined with 20 mg of the photocatalyst. Before sunlight exposure, the solutions were kept in the dark on magnetic stirrers for about 30 min to examine the contribution of absorption. And then, after a dark reaction, from the solutions took a 4 ml sample in a falcon tube with no sunlight at a pH of about 5–7. After this, the dye solutions were exposed to sunlight for 60 min with continuous magnetic stirring and every 15 min, samples of dye solutions were taken in a tube. After this, by using UV-visible spectroscopy, the absorbance of CR and RhB after each time interval was examined. The degradation of CR and RhB is observed to follow pseudo-first-order kinetics, which is:
(1)
where the first-order rate kinetics is represented by k and initial concentration and concentration at any time t is represented by and , respectively.
For each photocatalyst, the CR and RhB dye degradation percentages were calculated using the following equation:
(2)

Here, initial absorbance and final absorbance, respectively.

X-ray diffraction analysis

The X-ray diffraction patterns of Co0.5Sr0.5RExFe2−xO4, x = 0.00, and 0.06 are indicated in Figure 2(a)–2(c). For prepared spinel ferrites, the value of 2 from these patterns is observed in the range of 20°–70° to determine its lattice constant, average crystallite size, cell volume, lattice strain, micro-strain, and dislocation density. The patterns validated the presence of the prepared nanoparticles' cubic-spinel structure having a single phase and excluded out the possibility of any extra impurity peak. The reference card number 22-1086 was found to fit well with all of the diffraction peaks (220), (311), (400), (331), (422), (511), and (440). The classical Scherrer equation was applied to determine the average crystallite size (D) from the line width of the finest peaks:
(3)
Figure 1

Schematic representation for the synthesis of Co0.5Sr0.5RExFe2−xO4 (x = 0.00 and 0.06).

Figure 1

Schematic representation for the synthesis of Co0.5Sr0.5RExFe2−xO4 (x = 0.00 and 0.06).

Close modal
Figure 2

X-ray diffraction patterns of Co0.5Sr0.5Fe2O4 (a), Co0.5Sr0.5Nd0.06Fe1.94O4 (b), and Co0.5Sr0.5La0.06Fe1.94O4(c) spinel ferrites.

Figure 2

X-ray diffraction patterns of Co0.5Sr0.5Fe2O4 (a), Co0.5Sr0.5Nd0.06Fe1.94O4 (b), and Co0.5Sr0.5La0.06Fe1.94O4(c) spinel ferrites.

Close modal

Here, Scherer's constant is shown as k = 0.9, the wavelength is shown as , full-width half maximum (rad) is shown as and Brag's angle is given as (Kokare et al. 2018).

The following formula can be utilized to derive the lattice parameter ‘a’ and volume ‘V’:
(4)
(5)
where hkl denotes the miller indices values and ‘d’ denotes the gap between crystal planes.
The following formula is used to determine the lattice strain:
(6)

In this case, θ denotes the peak's highest diffraction angle.

Micro-strain can be calculated using the following relationship:
(7)
The following equation can be used to evaluate the dislocation density of the synthesized nanoparticles:
(8)

Here ‘D’ denotes for the crystalline size. Table 1 shows that the lattice parameter, cell volume, and average crystallite size all increase as lanthanum and neodymium ions are substituted. The findings for these structural properties are higher than those for pure Co–Sr spinel ferrite because of the formation of expansions and tensions inside the spinel structure. Additionally, the doping of La3+ and Nd3+ results in different values as a consequence of the higher ionic radii of lanthanum and neodymium ions over iron ions (Sharma et al. 2017). At the octahedral positions, La3+ and Nd3+ ions overlap with iron ions, which could have distorted the structure at the interstitial positions. As a result, the substitution of RE's metal ions may potentially affect structural properties like average crystallite size and lattice parameter. The values of lattice strain, micro-strain, and dislocation density decrease with the doping of lanthanum and neodymium, confirming that genuine spinel structure remains after the substitution of RE's metal ions (Yousaf et al. 2020).

Table 1

Variation of structural parameters with La3+ and Nd3+ doping concentration of Co0.5Sr0.5RExFe2−xO4 (x = 0.00 and 0.06) ferrite samples

ParametersCo0.5Sr0.5Fe2O4Co0.5Sr0.5 La0.06Fe1.94O4Co0.5Sr0.5 Nd0.06Fe1.94O4
Avg. crystalline size (nm) 31.51 36.59 36.60 
Lattice parameter (Å) 8.37 8.41 8.39 
Cell volume (a3586.63 594.82 592.50 
Dislocation density (lines/m) (10151.0066 0.7467 0.7464 
Lattice strain (10−33.6036 3.1208 3.113 
Micro-strain (lines−2/m−4) (10−30.5441 0.3785 0.414 
ParametersCo0.5Sr0.5Fe2O4Co0.5Sr0.5 La0.06Fe1.94O4Co0.5Sr0.5 Nd0.06Fe1.94O4
Avg. crystalline size (nm) 31.51 36.59 36.60 
Lattice parameter (Å) 8.37 8.41 8.39 
Cell volume (a3586.63 594.82 592.50 
Dislocation density (lines/m) (10151.0066 0.7467 0.7464 
Lattice strain (10−33.6036 3.1208 3.113 
Micro-strain (lines−2/m−4) (10−30.5441 0.3785 0.414 

Morphological analysis

Figure 3 shows the SEM micrographs of undoped and rare-earth's doped Co–Sr ferrite nanoparticles. The microstructures of the as-synthesized samples were revealed through SEM data. Particles are micro-structured and irregularly shaped according to the surface morphology. The boundaries of most of the particles are not clear in the SEM pictures, and many particles are agglomerated. As a result, average particle diameters of 0.4, 0.63, and 0.75 μm were obtained for Co0.5Sr0.5Fe2O4, Co0.5Sr0.5La0.06Fe1.94O4, and Co0.5Sr0.5Nd0.06Fe1.94O4 by using image J processing software and Origin Pro software, whereas the histograms of particle distribution of prepared spinel ferrites are shown in Figure 3. It has been reported that the average particle diameter rises when RE's metal ions are substituted in spinel ferrite (Gharagozlou 2009). In addition, grain size distribution was observed to be inhomogeneous.
Figure 3

SEM micrographs of pure and REs (La, Nd) doped Co–Sr spinel ferrite NPs, and SEM histograms for diameter distribution.

Figure 3

SEM micrographs of pure and REs (La, Nd) doped Co–Sr spinel ferrite NPs, and SEM histograms for diameter distribution.

Close modal

Optical analysis

The bandgap value of a semiconductor like spinel is important due to its optical and electrical properties in different applications, such as photocatalysis and photovoltaic devices. The optical characteristics, including bandgap, were shown to fluctuate with synthesis conditions and substitution (Das et al. 2010). Here, Tauc's relation is used to determine the optical bandgap:
(9)
where is the absorption co-efficient at a given wavelength , shows the energy bandgap, h is the plank's constant, is the frequency of light, A is a constant, and n represents the possible direct transition with a value of 1/2. Graphs of (h)2 versus energy (eV) are shown in Figure 4, where the linear portion is extrapolated to the h-axis to yield the energy bandgap (Nowsherwan et al. 2023). The bandgap values of the synthesized photocatalysts with and without La3+ and Nd3+ doping are also shown in Figure 4. The bandgap energy of pure Co0.5Sr0.5Fe2O4was determined to be 2.91 eV.Co0.5Sr0.5La0.06Fe1.94O4 and Co0.5Sr0.5Nd0.06Fe1.94O4 photocatalysts exhibited optical bandgaps around 2.62 and 2.52 eV, correspondingly. When compared to pure Co0.5Sr0.5Fe2O4, the bandgap energy of the La3+ and Nd3+ ions doped Co0.5Sr0.5RExFe2−xO4 nanoferrites was the lowest. Also, with the doping of La3+ and Nd3+ ions, the energy bandgaps reduced. And due to narrow bandgap value between 1.5 and 3 eV, ferrites are good for photocatalysis and have energy features that are beneficial for photocatalysis (López et al. 2015). So, these values of the energy bandgap make the photocatalysts highly efficient for dye degradation.
Figure 4

(a–c) The Tauc's plot of (αh2 versus energy (eV) for undoped and rare-earth ions (La3+, Nd3+) doped Co–Sr spinel ferrites.

Figure 4

(a–c) The Tauc's plot of (αh2 versus energy (eV) for undoped and rare-earth ions (La3+, Nd3+) doped Co–Sr spinel ferrites.

Close modal

Photocatalytic study

Various studies have verified the role of spinel ferrite and its nanocomposites in the purification of contaminated water. Thus, these nanoferrites are simple to regenerate and may be reutilized several times before losing their quality, resulting in lower treatment costs. These spinel ferrites are also useful for photocatalytic activity due to their narrow bandgap (<3 eV). When photons with higher energy are irradiated by sunlight, photo excitation occurs at the surface of the photocatalyst nanoparticles due to the equal number of holes and electrons created throughout the conduction and valence bands, resulting in the production of photogenerated electron–hole pairs. All this happens when the synthesized spinel ferrites have a bandgap energy value less than the energy of photons (Dong et al. 2015; Sun et al. 2019).
(10)
Depending on the oxidative environment and catalyst form, active free hydroxyl radicals are formed when H2O species from the surrounding environment react with photogenerated holes in the valence band and thus serve as an oxidant by degrading organic pollutants directly (Dong et al. 2015; Lassoued et al. 2018; Sun et al. 2019). In the same way, anionic super-oxide specie is made by electrons in the conduction band through reduction when these electrons react with adsorptive oxygen on nanoparticle surfaces. After this, hydroperoxyl radicals are made by the combination of anionic super-oxide specie with H+ or H2O, which then dissociate into H2O2 and O2 species (Li et al. 2011; Dong et al. 2015; Naik et al. 2019) Figure 5 represents the schematic illustration of photocatalytic degradation.
Figure 5

Photocatalytic degradation is depicted during sunlight.

Figure 5

Photocatalytic degradation is depicted during sunlight.

Close modal
The degradation of RhB and CR was tested on the synthesized undoped and rare-earth ions (La, Nd) doped Co–Sr spinel ferrites. The typical UV–vis spectra of CR and RhB dyes during degradation of the prepared spinel ferrites are shown in Figures 6 and 7. The generation and separation of electron–hole pairs, absorption, and diffusion are the abilities on which the degradation of water traces depends (Kelly et al. 1999; Ahire et al. 2019). The percentage of degradation increased as the UV exposure time increased; the detailed description is stated in Table 2. Also, Table 2 lists the degradation percentage for CR and RhB solutions that use all of the ferrites. From the table, it is obvious that with the substitution of RE ions, the degradation of CR and RhB increases. The dye degradation was enhanced due to the narrow bandgap of rare-earth ions doped in Co–Sr spinel ferrite in comparison to the undoped spinel ferrite. According to the literature, the photocatalytic activity increases with decreasing energy bandgap (Huixian et al. 2011).
Table 2

The degradation efficiency at various instants of time and pseudo-first-order rate constant for the photocatalytic degradation of cationic and anionic dyes

Ferrite compositionRhB t = 15RhB t = 30% Degradation
RhB
t = 45
RhB t = 60CR t = 15CR t = 30CR t = 45CR t = 60Rate constant ‘K’
RhB
CR
Co0.5Sr0.5Fe2O4 35 45 29 45 60 73 0.002 0.007 
Co0.5Sr0.5La0.06Fe1.94O4 50 57 61 67 27 50 77 81 0.003 0.016 
Co0.5Sr0.5Nd0.06Fe1.94O4 40 46 67 85 15 53 76 90 0.016 0.025 
Ferrite compositionRhB t = 15RhB t = 30% Degradation
RhB
t = 45
RhB t = 60CR t = 15CR t = 30CR t = 45CR t = 60Rate constant ‘K’
RhB
CR
Co0.5Sr0.5Fe2O4 35 45 29 45 60 73 0.002 0.007 
Co0.5Sr0.5La0.06Fe1.94O4 50 57 61 67 27 50 77 81 0.003 0.016 
Co0.5Sr0.5Nd0.06Fe1.94O4 40 46 67 85 15 53 76 90 0.016 0.025 
Table 3

Evaluation of photocatalytic degradation of present work with literature

MaterialsBandgap (Eg) (eV)DyeTime (min)% DegradationReference
Co0.7Mg0.3CexFe2−xO4 (x = 0.00 − 0.1) 1.53–1.57  MB 60 x = 0.00, 26.73%, x = 0.04, 74.94%, x = 0.1, 95.48% Basfer & Al-Harbi (2023)  
LaxMnFe2−xO4 (x = 0.0,0.04) 2.65–2.40 CV 150 x = 0.00, 32%, x = 0.04, 98% Baig et al. (2020)  
NiCo2O4(NCO),Ni0.95Ag0.05Co2O4(NACO), Ni0.95Ag0.05Co1.95La0.05O4 (NACLO) 1.7–1.9 EB 150 NCO = 80%, NACO = 88%, NACLO = 98% Priya et al. (2019)  
MSmxFe2–xO4,(M = Ni,Co; x = 0, 0.02, 0.06, 0.1) <2 MO, SO Ni (30 ∼ 100), Co (30 ∼ 100) Ni (MO = 93–94%, SO = 92–95%), Co (MO = 91–95%, SO = 93–96%) Singh et al. (2021)  
ZnAlFe1−xSmxO4 (x = 0, 0.02, 0.04, 0.06, 0.08) 2.1–2.2 EB 240 x = 0, 20%, x = 0.02, 43%, x = 0.04, 60%, x = 0.06, 78%, x = 0.08, 58%, Grecu et al. (2023)  
Co0.5Sr0.5RExFe2−xO4 (RE = La, Nd; x = 0, 0.06) 2.91–2.52 RhB, CR  60 Pure (RhB = 45%, CR = 73%), La (RhB = 67%, CR = 81%), Nd (RhB = 85%, CR = 90%) Present work 
MaterialsBandgap (Eg) (eV)DyeTime (min)% DegradationReference
Co0.7Mg0.3CexFe2−xO4 (x = 0.00 − 0.1) 1.53–1.57  MB 60 x = 0.00, 26.73%, x = 0.04, 74.94%, x = 0.1, 95.48% Basfer & Al-Harbi (2023)  
LaxMnFe2−xO4 (x = 0.0,0.04) 2.65–2.40 CV 150 x = 0.00, 32%, x = 0.04, 98% Baig et al. (2020)  
NiCo2O4(NCO),Ni0.95Ag0.05Co2O4(NACO), Ni0.95Ag0.05Co1.95La0.05O4 (NACLO) 1.7–1.9 EB 150 NCO = 80%, NACO = 88%, NACLO = 98% Priya et al. (2019)  
MSmxFe2–xO4,(M = Ni,Co; x = 0, 0.02, 0.06, 0.1) <2 MO, SO Ni (30 ∼ 100), Co (30 ∼ 100) Ni (MO = 93–94%, SO = 92–95%), Co (MO = 91–95%, SO = 93–96%) Singh et al. (2021)  
ZnAlFe1−xSmxO4 (x = 0, 0.02, 0.04, 0.06, 0.08) 2.1–2.2 EB 240 x = 0, 20%, x = 0.02, 43%, x = 0.04, 60%, x = 0.06, 78%, x = 0.08, 58%, Grecu et al. (2023)  
Co0.5Sr0.5RExFe2−xO4 (RE = La, Nd; x = 0, 0.06) 2.91–2.52 RhB, CR  60 Pure (RhB = 45%, CR = 73%), La (RhB = 67%, CR = 81%), Nd (RhB = 85%, CR = 90%) Present work 
Figure 6

Degradation profile of contaminant CR dye (UV–visible spectra) using Co0.5Sr0.5Fe2O4, Co0.5Sr0.5La0.06Fe1.94O4, and Co0.5Sr0.5Nd0.06Fe1.94O4 as photocatalysts.

Figure 6

Degradation profile of contaminant CR dye (UV–visible spectra) using Co0.5Sr0.5Fe2O4, Co0.5Sr0.5La0.06Fe1.94O4, and Co0.5Sr0.5Nd0.06Fe1.94O4 as photocatalysts.

Close modal
Figure 7

The change in absorption of RhB solution with time in the presence of Co0.5Sr0.5Fe2O4, Co0.5Sr0.5La0.06Fe1.94O4, and Co0.5Sr0.5Nd0.06Fe1.94O4 photocatalysts under solar irradiation.

Figure 7

The change in absorption of RhB solution with time in the presence of Co0.5Sr0.5Fe2O4, Co0.5Sr0.5La0.06Fe1.94O4, and Co0.5Sr0.5Nd0.06Fe1.94O4 photocatalysts under solar irradiation.

Close modal
Figure 8(a) and 8(b) display the photocatalytic degradation of RhB and CR dyes, and degradation efficiency increases with time, reaching a maximum between 30 and 60 min for the rare-earth substituted Co–Sr spinel ferrites. As a result of this doping, defects and surface oxygen vacancies are formed, promoting photocatalytic degradation.
Figure 8

C0/Ct vs time for CR (a) and RhB (b) degradation utilizing CO0.5Sr0.5RExFe2−xO4 nanoferrites (x = 0.00, and 0.06).

Figure 8

C0/Ct vs time for CR (a) and RhB (b) degradation utilizing CO0.5Sr0.5RExFe2−xO4 nanoferrites (x = 0.00, and 0.06).

Close modal
The as-synthesized samples' ln(C0/Ct) against time plot is shown in Figure 9(a) and 9(b). The photodegradation of CR and RhB is often observed to adopt pseudo-first-order kinetics. The highest values of rate constant were observed to be 0.016 and 0.025 min−1 for Co0.5Sr0.5Nd0.06Fe1.94O4 for degrading RhB and CR according to Table 2. With increasing reaction time, the degradation rate also increased and the highest degradation occurred between 30 and 60 min. The photogenerated electrons released in the catalytic process exhibit significant deoxidizing and oxidizing characteristics due to lanthanum and neodymium doping in the Co–Sr ferrite lattice, and then maximum degradation of CR and RhB dyes is achieved. At the lattice's surface, as a result of a rise in defects and surface oxygen vacancies, this can happen. Another possibility is that metal ions are distributed differently at the octahedral and tetrahedral locations, affecting catalytic efficiency in the spinel ferrite as in A and B sites. As the catalytic activity is primarily dependent on metal ions in the B position, the catalytic efficiency for metal ions in the A position is often lower. At B sites, the reactant molecules interact more easily with metal ions because, at these sites, the metal ions are separated by a significant distance. When ‘La3+’ and ‘Nd3+’ ions are inserted separately into the Co–Sr ferrite lattice, these ions appear in the B positions, resulting in high catalytic efficiency due to the large concentration of lanthanum and neodymium ions (Ali et al. 2020).
Figure 9

(a,b) Pseudo-first-order kinetics of prepared photocatalysts.

Figure 9

(a,b) Pseudo-first-order kinetics of prepared photocatalysts.

Close modal
Figure 10(a) and 10(b) show that the degradation of RhB and CR dyes increases over time for the prepared samples of ‘La3+’- and ‘Nd3+’-substituted Co–Sr spinel ferrites. The degradation efficiency improved over time, with significant degradation achieved in an hour for the obtained samples, as shown in Figure 10(a) and 10(b), respectively. Moreover, using the photocatalytic mechanism to degrade CR and RhB is a safe procedure. As the lanthanum and neodymium substituted Co–Sr spinel ferrites pose no environmental risk, water purification can be effectively done by using them.
Figure 10

Percent degradation efficiency of CR (a) and RhB (b) through Co0.5Sr0.5RExFe2−xO4 (x = 0.00, and 0.06) nanoparticles.

Figure 10

Percent degradation efficiency of CR (a) and RhB (b) through Co0.5Sr0.5RExFe2−xO4 (x = 0.00, and 0.06) nanoparticles.

Close modal

Pure and rare-earth metal ions (La, Nd)-substituted Co0.5Sr0.5RExFe2−xO4 (x = 0.00, and 0.06) nanoferrites were prepared using an inexpensive sol–gel approach. Powder XRD analysis demonstrates the emergence of single-phase face-centered cubic-spinel structures. Since REs have wider ionic radii than iron elements, the substitution of these cations alters the lattice parameter and crystalline size. By substituting REs3+, the average crystallite size and optical energy bandgap values varied from 31 to 36 nm and from 2.91 to 2.52 eV, respectively. SEM analysis allowed for the identification of the microstructures of the synthesized ferrites. Inhomogeneous grains with average particle diameters ranging from 0.4 to 0.75m were seen in the SEM pictures of the prepared spinel ferrites. The substitution of La3+ and Nd3+ ions enhanced the degradation activity of the Co–Sr spinel ferrite samples. In this study, we achieved a significant improvement in the photocatalytic activity of CR and RhB dyes, attaining 90 and 85%, respectively. Amazingly, the nano-ferrite sample Co0.5Sr0.5Nd0.06Fe1.94O4 exhibited the highest percentages of CR and RhB degradation over the duration of 60 min because it had the lowest energy gap, and it can be used in treatments to clean up environmental pollutants effectively (Table 3).

This work is acknowledged by HEC-funded project NRPU#10408.

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Ahalya
K.
,
Suriyanarayanan
N.
&
Ranjithkumar
V.
2014
Effect of cobalt substitution on structural and magnetic properties and chromium adsorption of manganese ferrite nano particles
.
Journal of Magnetism and Magnetic Materials
372
,
208
213
.
Ahire
S. A.
,
Bachhav
A. A.
,
Bagul
V. R.
,
Ahire
B. B.
,
Tupe
U. J.
&
Nashik
M. I.
2019
A literature review on: synthesis approaches of metal doped titanium dioxide
6
,
1238
1250
.
Ali
N.
,
Said
A.
,
Ali
F.
,
Raziq
F.
,
Ali
Z.
,
Bilal
M.
,
Reinert
L.
,
Begum
T.
&
Iqbal
H. M.
2020
Photocatalytic degradation of Congo red dye from aqueous environment using cobalt ferrite nanostructures: development, characterization, and photocatalytic performance
.
Water, Air, & Soil Pollution
231
,
1
16
.
Baig
M. M.
,
Zulfiqar
S.
,
Yousuf
M. A.
,
Touqeer
M.
,
Ullah
S.
,
Agboola
P.
,
Warsi
M. F.
&
Shakir
I.
2020
Structural and photocatalytic properties of new rare earth La3+ substituted MnFe2O4 ferrite nanoparticles
.
Ceramics International
46
,
23208
23217
.
Bensebaa
F.
,
Zavaliche
F.
,
L'ecuyer
P.
,
Cochrane
R.
&
Veres
T.
2004
Microwave synthesis and characterization of Co–ferrite nanoparticles
.
Journal of Colloid and Interface Science
277
,
104
110
.
Chahar
D.
,
Taneja
S.
,
Bisht
S.
,
Kesarwani
S.
,
Thakur
P.
,
Thakur
A.
&
Sharma
P.
2021
Photocatalytic activity of cobalt substituted zinc ferrite for the degradation of methylene blue dye under visible light irradiation
.
Journal of Alloys and Compounds
851
,
156878
.
Das
N.
,
Ghosh
P.
,
Mitra
M.
&
Chattopadhyay
K.
2010
Effect of film thickness on the energy band gap of nanocrystalline CdS thin films analyzed by spectroscopic ellipsometry
.
Physica E: Low-Dimensional Systems and Nanostructures
42
,
2097
2102
.
Deraz
N.
&
Hessien
M.
2009
Structural and magnetic properties of pure and doped nanocrystalline cadmium ferrite
.
Journal of Alloys and Compounds
475
,
832
839
.
Dhiman
M.
&
Singhal
S.
2019
Enhanced catalytic properties of rare-earth substituted cobalt ferrites fabricated by sol-gel auto-combustion route
.
Materials Today: Proceedings
14
,
435
444
.
Di Mauro
A.
,
Fragala
M. E.
,
Privitera
V.
&
Impellizzeri
G.
2017
ZnO for application in photocatalysis: from thin films to nanostructures
.
Materials Science in Semiconductor Processing
69
,
44
51
.
Dong
S.
,
Feng
J.
,
Fan
M.
,
Pi
Y.
,
Hu
L.
,
Han
X.
,
Liu
M.
,
Sun
J.
&
Sun
J.
2015
Recent developments in heterogeneous photocatalytic water treatment using visible light-responsive photocatalysts: a review
.
Rsc Advances
5
,
14610
14630
.
Gao
F.
,
Chen
X.
,
Yin
K.
,
Dong
S.
,
Ren
Z.
,
Yuan
F.
,
Yu
T.
,
Zou
Z.
&
Liu
J. M.
2007
Visible-light photocatalytic properties of weak magnetic BiFeO3 nanoparticles
.
Advanced Materials
19
,
2889
2892
.
Gao
X.
,
Liu
L.
,
Birajdar
B.
,
Ziese
M.
,
Lee
W.
,
Alexe
M.
&
Hesse
D.
2009
High-density periodically ordered magnetic cobalt ferrite nanodot arrays by template-assisted pulsed laser deposition
.
Advanced Functional Materials
19
,
3450
3455
.
Gao
T.
,
Chen
Z.
,
Zhu
Y.
,
Niu
F.
,
Huang
Q.
,
Qin
L.
,
Sun
X.
&
Huang
Y.
2014
Synthesis of BiFeO3 nanoparticles for the visible-light induced photocatalytic property
.
Materials Research Bulletin
59
,
6
12
.
Goldman
A.
2006
Modern Ferrite Technology
.
Springer Science & Business Media
, New York, USA.
Grecu
I.
,
Samoila
P.
,
Pascariu
P.
,
Cojocaru
C.
,
Ignat
M.
,
Dascalu
I.-A.
&
Harabagiu
V.
2023
Enhanced photodegradation of organic pollutants by novel samarium-doped zinc aluminium spinel ferrites
.
Catalysts
13
,
266
.
Ishaque
M.
,
Islam
M.
,
Khan
M. A.
,
Rahman
I.
,
Genson
A.
&
Hampshire
S.
2010
Structural, electrical and dielectric properties of yttrium substituted nickel ferrites
.
Physica B: Condensed Matter
405
,
1532
1540
.
Kelly
J.
,
Kooij
E. S.
&
Meulenkamp
E.
1999
Luminescence studies of semiconductor electrodes
.
Electrochimica Acta
45
,
561
574
.
Lassoued
A.
,
Lassoued
M. S.
,
Dkhil
B.
,
Ammar
S.
&
Gadri
A.
2018
Improved photocatalytic activities of CuxCo0.5-xNi0.5Fe2O4 nanoparticles through co-precipitation method in degrading methylene blue
.
Physica E: Low-Dimensional Systems and Nanostructures
101
,
29
37
.
Liu
Z.
,
Qi
Y.
&
Lu
C.
2010
High efficient ultraviolet photocatalytic activity of BiFeO3 nanoparticles synthesized by a chemical coprecipitation process
.
Journal of Materials Science: Materials in Electronics
21
,
380
384
.
López
Y. O.
,
Vázquez
H. M.
,
Gutiérrez
J. S.
,
Velderrain
V. G.
,
Ortiz
A. L.
&
Martínez
V. C.
2015
Synthesis method effect of CoFe2O4 on its photocatalytic properties for H2 production from water and visible light
.
Journal of Nanomaterials
16
,
76
76
.
Maaz
K.
,
Karim
S.
,
Mumtaz
A.
,
Hasanain
S.
,
Liu
J.
&
Duan
J.
2009
Synthesis and magnetic characterization of nickel ferrite nanoparticles prepared by co-precipitation route
.
Journal of Magnetism and Magnetic Materials
321
,
1838
1842
.
Melagiriyappa
E.
&
Jayanna
H.
2009
Structural and magnetic susceptibility studies of samarium substituted magnesium–zinc ferrites
.
Journal of Alloys and Compounds
482
,
147
150
.
Mironyuk
I.
,
Tatarchuk
T.
,
Naushad
M.
,
Vasylyeva
H.
&
Mykytyn
I.
2019a
Highly efficient adsorption of strontium ions by carbonated mesoporous TiO2
.
Journal of Molecular Liquids
285
,
742
753
.
Mironyuk
I.
,
Tatarchuk
T.
,
Vasylyeva
H.
,
Gun'ko
V. M.
&
Mykytyn
I.
2019b
Effects of chemosorbed arsenate groups on the mesoporous titania morphology and enhanced adsorption properties towards Sr (II) cations
.
Journal of Molecular Liquids
282
,
587
597
.
Naik
M. M.
,
Naik
H. B.
,
Nagaraju
G.
,
Vinuth
M.
,
Vinu
K.
&
Viswanath
R.
2019
Green synthesis of zinc doped cobalt ferrite nanoparticles: structural, optical, photocatalytic and antibacterial studies
.
Nano-Structures & Nano-Objects
19
,
100322
.
Naik
M. M.
,
Vinuth
M.
,
Karthik
K.
,
Suresha
B.
,
Nagaraju
G.
&
Sujatha
H.
2020
Photocatalytic degradation of dyes by cobalt ferrite nanoparticles synthesized by sol-gel method
. In
AIP Conference Proceedings
.
AIP Publishing LLC
, p.
040004
.
Niu
F.
,
Gao
T.
,
Zhang
N.
,
Chen
Z.
,
Huang
Q.
,
Qin
L.
,
Sun
X.
&
Huang
Y.
2015b
Hydrothermal synthesis of bifeo3 nanoparticles for visible light photocatalytic applications
.
Journal of Nanoscience and Nanotechnology
15
,
9693
9698
.
Nowsherwan
G. A.
,
Zaib
A.
,
Shah
A. A.
,
Khan
M.
,
Shakoor
A.
,
Bukhari
S. N. S.
,
Riaz
M.
,
Hussain
S. S.
,
Shar
M. A.
&
Alhazaa
A.
2023
Preparation and numerical optimization of TiO2: CdS thin films in double perovskite solar cell
.
Energies
16
,
900
.
Ponraj
C.
,
Vinitha
G.
&
Daniel
J.
2017
A review on the visible light active bifeo3 nanostructures as suitable photocatalyst in the degradation of different textile dyes
.
Environmental Nanotechnology, Monitoring & Management
7
,
110
120
.
Priya
A. S.
,
Geetha
D.
,
Karthik
K.
&
Rajamoorthy
M.
2019
Investigations on the enhanced photocatalytic activity of (Ag, La) substituted nickel cobaltite spinels
.
Solid State Sciences
98
,
105992
.
Rana
M. U.
&
Abbas
T.
2002
The effect of Zn substitution on microstructure and magnetic properties of Cu1− xZnxFe2O4 ferrite
.
Journal of Magnetism and Magnetic Materials
246
,
110
114
.
Rezlescu
E.
,
Sachelarie
L.
,
Popa
P.
&
Rezlescu
N.
2000
Effect of substitution of divalent ions on the electrical and magnetic properties of Ni-Zn-Me ferrites
.
IEEE Transactions on Magnetics
36
,
3962
3967
.
Schneider
J.
,
Matsuoka
M.
,
Takeuchi
M.
,
Zhang
J.
,
Horiuchi
Y.
,
Anpo
M.
&
Bahnemann
D. W.
2014
Understanding TiO2 photocatalysis: mechanisms and materials
.
Chemical Reviews
114
,
9919
9986
.
Sharma
R.
,
Thakur
P.
,
Sharma
P.
&
Sharma
V.
2017
Ferrimagnetic Ni2+ doped Mg-Zn spinel ferrite nanoparticles for high density information storage
.
Journal of Alloys and Compounds
704
,
7
17
.
Tavakoli-Azar
T.
,
Mahjoub
A. R.
,
Seyed Sadjadi
M.
,
Farhadyar
N.
&
Hossaini Sadr
M.
2020
Synthesis and characterization of CdTiO3@ S composite: investigation of photocatalytic activity for the degradation of crystal violet under sun light
.
Journal of Inorganic and Organometallic Polymers and Materials
30
,
667
683
.
Verdaguer
D.
,
Jansen
M. A.
,
Llorens
L.
,
Morales
L. O.
&
Neugart
S.
2017
UV-A radiation effects on higher plants: exploring the known unknown
.
Plant Science
255
,
72
81
.
Vosoughifar
M.
&
Kimiay
A.
2016
Neodymium-doped copper ferrite: auto-combustion synthesis, characterization and photocatalytic properties
.
Journal of Materials Science: Materials in Electronics
27
,
10031
10035
.
Yu
K.
,
Yang
S.
,
He
H.
,
Sun
C.
,
Gu
C.
&
Ju
Y.
2009
Visible light-driven photocatalytic degradation of rhodamine B over nabio3: pathways and mechanism
.
The Journal of Physical Chemistry A
113
,
10024
10032
.

Author notes

Equal authors.

This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY-NC-ND 4.0), which permits copying and redistribution for non-commercial purposes with no derivatives, provided the original work is properly cited (http://creativecommons.org/licenses/by-nc-nd/4.0/).