Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/364298094

Synthesis of spinel ferrites and application in the photodegradation of


Rhodamine in water via pH control assistance

Article in Optical Materials · October 2022


DOI: 10.1016/j.optmat.2022.113043

CITATIONS READS

0 100

3 authors:

Fabiano Praxedes Marcos Augusto Lima Nobre


São Paulo State University São Paulo State University
14 PUBLICATIONS 60 CITATIONS 223 PUBLICATIONS 2,317 CITATIONS

SEE PROFILE SEE PROFILE

Silvania Lanfredi
São Paulo State University
161 PUBLICATIONS 2,338 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Updates of Impact Factors (SCI, SSCI, EI, F1000, etc) View project

Synthesis and characterization of nanostructured materials applied in electro-electronic devices View project

All content following this page was uploaded by Marcos Augusto Lima Nobre on 10 October 2022.

The user has requested enhancement of the downloaded file.


Optical Materials xxx (xxxx) 113043

Contents lists available at ScienceDirect

Optical Materials
journal homepage: www.elsevier.com/locate/optmat

Research Article

F
Synthesis of spinel ferrites and application in the photodegradation of
Rhodamine in water via pH control assistance

OO
M.C.M. Ferreira a, c, F.R. Praxedes a, c, M.A.L. Nobre b, c, S. Lanfredi a, c, *
a Department of Chemistry and Biochemistry, São Paulo State University (Unesp), School of Technology and Sciences, R. Roberto Simonsen – 305, CEP: 19060–900,
Presidente Prudente- SP, Brazil
b Department of Physics, São Paulo State University (Unesp), School of Technology and Sciences, R. Roberto Simonsen – 305, CEP: 19060–900, Presidente Prudente- SP,

Brazil
c São Paulo State University (Unesp), Institute of Biosciences, Humanities and Exact Sciences, São José do Rio Preto, SP, 15054–000, Brazil

ARTICLE INFO

Keywords:
Ferrites
Spinel structure
Rhodamine 6G
ABSTRACT

PR
Nanostructured powders of cobalt ferrite with stoichiometry CoxFe3-xO 4, where 0.4 ≤ x ≤ 1.2, were synthesized
by the wet chemical route from the Co-precipitation method. Single-phase and crystalline nanopowders of spinel
type structure with cubic symmetry and space group were characterized by X-ray diffraction. The struc-
tural refinement was carried out by the Rietveld Method showing a transition of normal spinel to an inverse
D
Photocatalysis
spinel, as a function of Co2+concentration. Both spectra, Infrared Spectroscopy and Diffuse reflectance spec-
troscopy were used to characterize chemical bonds Co2+ in tetrahedral coordination, for higher cobalt concentra-
tion, and band gap analysis, respectively. The photoactivity of powders was evaluated from the photodegrada-
TE

tion of the Rhodamine 6G dye under UV light. CoFe2O4 nanoparticles showed the best photocatalytic degradation
with 90.9% in an alkaline medium, which obeys a pseudo-first order kinetic reaction, being the maximum appar-
ent rate constant equal to 10.23 × 10−3. Effects of the occupation of tetrahedral and octahedral sites in the spinel
structure in the photoactivity parameter are discussed.
EC

1. Introduction absorption and catalytic properties, binary ferrites compositions have


been proposed via the partial substitution of iron (III) by metal cations
Advanced oxidation processes as heterogeneous photocatalysis have as Co2+, Zn2+, and Ni2+ in the spinel structure [5–7,9]. Cobalt ferrite
received special attention for environmental and industrial application, CoFe2O4 is a classical n-type semiconductor with a narrow type band
as well as for clean energy generation [1,2]. Semiconductors, with mul- gap of 1.76 eV [10]. The magnetic property of ferrite-based materials
tifunctional properties, are potential photocatalysts due to the coupled has scored these materials as promising photocatalysts due to the possi-
principle of solar photon absorption [2–5]. Spinel ferrite- bility of easy recuperation from the aqueous medium by applying a
RR

semiconductor is a class of multifunctional materials with intrinsic magnetic field [11–14]. However, some aspects related to the photoac-
magnetic feature, which combined with the solar photo-absorption and tivity of cobalt ferrite with A:B different from 1:2 and its structure is
the adsorbent properties [5–8], enables its use in environmental reme- still unclear; since cobalt ferrite can be stabilized in two distinct cubic
diation with easy retrieval of photocatalyst from the medium by using structures called spinel and spinel inverse [15].
an applied magnetic field. In a broad sense, if A = B = Fe, the cobalt ferrite is transformed in
In recent years, ferrites with spinel structure have been investigated Fe3O4. As a matter of fact, the ferrite Fe3O4 can be further considered a
CO

for their photocatalytic properties [8]. The spinel structure has been prototype of other new functional or polyfunctional phases. In this
represented by formula AB2O4, where A and B represent sites of metal- study, binary cobalt ferrite nanoparticles with stoichiometry CoxFe3-xO4
lic cation exhibiting tetrahedral octahedral coordination with oxygen (0.4 ≤ x ≤ 1.2) were prepared by the wet chemical, from the Co-
atoms, respectively. The small band gap-covering photon absorption at precipitation method. The Co-precipitation was chosen due to several
visible light region of electromagnetic spectra makes A:B with a ratio of advantages, such as facile and rapid preparation, cost-effectiveness, and
1:2 mono-ferrite a poor photocatalyst. In order to improve the photo- high mass production [16]. Detailed structural characterization was

* Corresponding author. Department of Chemistry and Biochemistry, São Paulo State University (Unesp), School of Technology and Sciences, R. Roberto Simonsen

– 305, CEP: 19060–900, Presidente Prudente- SP, Brazil.


E-mail address: silvania.lanfredi@unesp.br (S. Lanfredi).

https://doi.org/10.1016/j.optmat.2022.113043
Received 30 May 2022; Received in revised form 6 September 2022; Accepted 22 September 2022
0925-3467/© 20XX

Note: Low-resolution images were used to create this PDF. The original images will be used in the final composition.
M.C.M. Ferreira et al. Optical Materials xxx (xxxx) 113043

carried out and the photocatalytic activity on the structural and physic- with sphere integration ISR-2600Plus model equipped with a diffuse re-
ochemical properties of cobalt ferrites was established. flectance module.

2. Experimental procedure 2.3. Adsorption and photocatalytic tests

2.1. Synthesis of cobalt ferrite by the wet chemical reaction The photocatalytic activity of the CFO-0.4, CFO-0.8, CFO-1.0, and
CFO-1.2 samples was evaluated following the kinetics of degradation of
A set of cobalt doped ferrite powders with formula Co0.4Fe2.6O4 Rhodamine 6G (C28H31N2O3Cl, 99%, Sigma-Aldrich) called Rhd6G. Ex-
(CFO-0.4), Co0.8Fe2.2O4 (CFO-0.8), CoFe2O4 (CFO-1.0) and Co1.2Fe1.8O4 periments were carried out in polyvinyl chloride (PVC) homemade re-
(CFO-1.2) were synthesized via wet chemical using the Co-precipitation actor (250 mL) equipped with a mercury lamp (Osram, 15 W, maxi-

F
method. Precursor solutions of the CFO-0.4, CFO-0.8, CFO-1.0 and mum emission at ca. 254 nm) and a water pump (5W) operating at a
CFO-1.2 were prepared using cobalt nitrate, Co(NO3)2·6H2O (98%, flow rate of 180 L h−1 [22]. The total volume of Rhd6G solution was
Synth), and iron nitrate, Fe(NO3)3·9H2O (98%, Synth) in deionized wa- 1 L, while the volume of irradiated solution in the photoreactor was ca.

OO
ter. The solution prepared, with concentration equal to 0.14 mol L−1, 250 mL. The initial concentration of Rhd6G and the amount of the cata-
was heated at 80 °C under continuous stirring. Cobalt ferrites were ob- lysts used in all photocatalytic tests were 10.4 μmol L−1 and 100 mg, re-
tained via alkali precipitation with NaOH solution (2 mol L−1). A spectively. Therefore, the loading of samples was maintained keep con-
change in the solution's color from brown to dark was observed when stant at ca. 0.1 g L−1. The equilibrium of adsorption of Rhd6G in the
the precipitates were formed. The pH of the solution was equal to 9.0. dark was achieved after 1 h under constant stirring. After this time, UV
The system was kept under heating and stirring for 2 h until complete irradiation was performed for the 4 h of reaction. Several aliquots were
precipitation of the powders. According to Zhao et al. [17], the synthe- taken each 15 min for the first hour of irradiation and each 30 min for

PR
sis of cobalt ferrite by the Co-precipitation method can be represented the next 3 h. Aliquots were centrifuged at 3000 rpm before analysis by
by the following equations: UV–Vis spectroscopy in a Genesis model spectrophotometer with a sin-
gle bean. The samples were measured at 525 nm as the wavelength of
Co(NO3)2·6H2O → Co(H2O)62+ + 2NO3− (1) maximal absorption for Rhodamine Rhd6G dye. Experiments were car-
ried out at neutral (7.0), acid (3.0), and alkaline pH (9.0). Thus, the in-
Fe(NO3)3·9H2O → Fe(H2O)63+ + 3NO3− + 3H2O (2)
fluence of the pH of the solution upon adsorption and photocatalytic ac-
Co(H2O)62+ + NaOH → Co(OH)x2−x → Co(OH)2↓ → Co(OH)3− (3) tivity was verified. For the tests in neutral and alkaline medium, the so-
lution was adjusted to 7.0, or 9.0 using NaOH solution ca. 2 mol L−1,
D
Fe(H2O)63+ + NaOH → Fe(OH)y3−y → Fe(OH)3↓ → Fe(OH)4− (4) while for the experiments under acidic conditions were not necessary to
(5) adjust the pH.
Co(OH)3− + 2Fe(OH)4− → CoFe2O4↓ + 4H2O + 3OH−

3. Results and discussion


TE

The hydroxy metal complex (reaction 5) is the main path to gener-


ate the cobalt ferrite particles by the Co-precipitation method.
The precipitate was centrifuged (6000 rpm for 40 min) and washed 3.1. Morphological analysis of cobalt ferrite
thrice with deionized water and ethanol. Then, precipitates were dried
at 80 °C for 24h. The set of SEM micrographs of the cobalt ferrite powders are shown
in Fig. 1a–d. The results indicate that the samples prepared using the
EC

2.2. Characterization Co-precipitation method are nanoparticulate in a spherical shape. How-


ever, the cobalt ferrite particles present a tendency to form clusters, as
The morphology of the particles was investigated by Scanning Elec- shown in the scheme in Fig. 1e. According to Zhao et al. [17], the syn-
tron Microscopy (SEM) using Carl Zeiss equipment, model EVO LS15. thesis of cobalt ferrite by the Co-precipitation method follows equations
The structural characterization of Co-doped ferrite powders was an- (1)–(5)(see section 2.1). The average particle size parameter decreases
alyzed by X-ray diffraction (XRD). The patterns were recorded in the re- as a function of Co amount in crystalline lattice, as observed in the his-
tograms of Fig. 1a–d. The particle size was estimated using Euler's theo-
RR

gion of 5o ≤ 2θ ≤ 80o with a 0.02° scanning step using a Shimadzu dif-


fractometer (XRD-6000 model) with Cu-Kα radiation (λ = 1.54 Å) and rem [22], where the average particle size <P> is obtained from the
a monochromator of graphite. Crystalline structures of Co-doped fer- particle size distribution ρ(n) by means of the n range of the particles,
rites powders were refined by the Rietveld method using FullProf soft- according to equation (6):
ware [18,19]. The set of parameters and variables refined are summa-
<P ≥ Σn.ρ(n) (6)
rized, as follows: background coefficients using a polynomial of 5-order
with a pseudo-Voigt function; scale factor; lattice parameter; and The average particle size of cobalt ferrite from the histograms in Fig.
CO

atomic coordination for three atoms (Fe, Co, O). The crystalline struc- 1a–d were ca. 133.5 nm, 132 nm, 74.5 nm, and 102.3 nm for CFO-0.4,
tures were built using Vesta software [20] from atomic positions de- CFO-0.8, CFO-1.0, and CFO-1.2, respectively. In addition, the particle
rived by the refinement. clusters formation is affected with a tendency to be less agglomerated,
The chemical bonds of the powders were analyzed by Fourier- as shown in Fig. 1a–d and schematized in Fig. 1e. Thus, considering the
transform infrared spectroscopy (FTIR) using a PerkinElmer spec- experimental conditions studied here, NaOH solution (2 mol L−1) drop-
trophotometer (Frontier model), equipped with a diamond-crystal for wise increase rapidly the local concentration in the cationic precursor
the attenuated total reflection (ATR). The measurements were carried solution (0.14 mol L−1). Momentaneous supersaturation produces the
out with an instrument resolution of 2 cm−1 in the range of primary particles or nuclei, which Ostwald-ripening governs the crystal
4000–250 cm− 1 for 60 scans. growth. The Ostwald ripening is schematized in Fig. 1f.
The energy band gap of the samples was determined from the The mass of these primary particles is available to grow larger parti-
Kubelka-Munk theory based on the Tauc method. The band gap was de- cles and coarsen the size distribution [23,24]. Therefore, a metastable
termined by plotting (F(R)hν)1/α = B (hν – Eg) considering α = 1/2 for supersaturated phase will not only generate primarily particles [23] but
a direct transition [21]. The UV–vis analysis was carried out in the will also cause deposition of the generated primary particles and conse-
range of 220–1400 nm using a Shimadzu UV2600 spectrophotometer quent growth of the clusters [24]. The growth and nucleation paths fol-
low the Gibbs–Thomson effect, where smaller clusters dissolve and

2
M.C.M. Ferreira et al. Optical Materials xxx (xxxx) 113043

F
OO
PR
D
TE

Fig. 1. SEM micrographs of the powders of: a) CFO-0.4, b) CFO-0.8, c) CFO-1.0, and d) CFO-1.2. e) Schematic representation of cobalt ferrite particles pro-
duced via wet chemical, according to cobalt-doping. f) Schematic representation of the Ostwald-ripening process in the formation of cobalt ferrite particle.

transfer their mass to larger clusters. Therefore, the increase in cobalt- ing to Fig. 2d, the spinel structure is arranged in a close-packed cubic
doping can influence the thermodynamical stability and/or modifying lattice formed by oxygen atoms (red sphere) and with the iron and
EC

mass transport mechanism of primary particles, generating more stable cobalt ions distributed at two different crystallographic sites, tetrahe-
smaller particles and clusters in CFO-1.0 and CFO-1.2 than in CFO-0.4 dral A-site (pink sphere) and octahedral B-site (purple sphere). The
and CFO-0.8. crystallite anisotropic shape of cobalt ferrite powders is shown in Fig.
2e, and the average crystallite size (D) is listed in Table 1. The crystal-
3.2. Structural analysis lite shape, size and symmetry are influenced by cobalt-doping, assum-
ing a more cubic shape at the higher Co2+ concentration. The cubic
RR

Fig. 2a shows diffractograms of CFO-0.4, CFO-0.8 CFO-1.0 and crystallite shape agrees with the TEM data reported by Karthickraja et
CFO-1.2 powders synthesized using the Co-precipitation method. Dif- al. [29], Maaz et al. [30], and Zhang et al. [31]. In addition, the average
fraction patterns of all powders were assigned to spinel structure with crystallite size (Table 1) increases from 4.6 to 8.8 nm in the range of
cubic symmetry and space group (227) according to JCPDS card CFO-0.4 to CFO-1.0 and decreases to 3.9 nm for the CFO-1.2. As men-
number #22–1086 [25]. Diffraction lines of another phases are not evi- tioned in the SEM discussion, the crystallization of cobalt ferrite is gov-
denced, what confirms the Co2+-doping in the spinel structure and the erned by the Gibbs–Thomson effect [23,24]. In this sense, the growth
reaching of single-phase powders. and nucleation paths of cobalt ferrite crystallization change according
CO

Fig. 2b shows that the line hkl (311) exhibits a shifting of line of ca. to cobalt-doping, where at low Co2+ concentration, the mass transfer is
0.23° and an increase in the intensity as cobalt-doping increases. The favorable, allowing the development of larger crystallites. Furthermore,
shifting of diffraction line to a high degree can be associated with a the concentration of Co2+ higher than CFO-1.0 has an inverse effect re-
shortening of the interatomic distance and some degree of uniform ten- stricting the crystal growth. Therefore, the cobalt diffusion process in
sion [26]. The tensile strain due to an internal compressive micro-stress the spinel structure is thermodynamically favorable up to the CFO-1.0
[27] causes a decreasing in unit cell volume as discussed further. How- composition. According to Muhich et al. [32], cobalt and iron have sim-
ever, an increase in the intensity of the cobalt-doping can be related to ilar activation energies when they diffuse with +2 oxidation state. Ac-
a change in the atomic positions in the spinel structure with cubic sym- cordingly, the range of CFO-0.4 to CFO-1.0 exhibits an increase in crys-
metry [26]. tallite size from 4.6 to 8.8 nm, as shown in Table 1. However, the excess
In order to investigate the structural correlation with cobalt-doping, of cobalt-doped in the CFO-1.2 ferrite hampers the growth of crystal-
the structural refinement was performed using the Rietveld method lites (3.9 nm). The main reason is the difference in activation energy in
[22]. In addition, the crystallite size and anisotropic shape were mod- the diffusion of cobalt and iron with +2 and + 3 oxidation states, re-
eled by the Rietveld refinement using the spherical harmonic approach spectively [32]. In this sense, the diffusion of Co2+ governs the crystal-
[22,28]. Fig. 2c shows the structural refinement of CFO-1.0 as a model, lite growth mechanism in cobalt ferrites. Furthermore, smaller crystal-
and the crystalline structure representation is shown in Fig. 2d. Accord- lites benefit the formation of clusters, as in CFO-0.4 as in CFO-1.2.

3
M.C.M. Ferreira et al. Optical Materials xxx (xxxx) 113043

F
OO
PR
D
TE
EC

Fig. 2. a X-ray diffraction of cobalt ferrite powders. b Expansion of the 2θ region between 34° and 38°. c Structural refinement by the Rietveld method for CFO-1.0
as a model, with cubic symmetry (227). d Spinel structure representation of cobalt ferrite, where pink tetrahedron correspond to atomic A-sites and purple oc-
tahedron are related to atomic B-sites. e Crystallite anisotropic shape of cobalt ferrite samples modeled from spherical harmonic approach via Rietveld refinement.
RR

Table 1 tutional solid solution [26,33]. The non-isovalent substitution and the
Lattice parameters and unit cell volumes for the cobalt ferrite powders. difference in ionic radii between Co2+ (0.78 Å) and Fe3+ (0.67 Å) can
Powder Lattice Parameter Unit cell Volume Rp Rwp Rexp χ2 D lead to the lattice distortion in the cubic spinel structure of the ferrite
(Å) (Å3) (nm) [26,27,33,34]. Thus, the correlation between structure distortion and
atomic substitution can be related to the bond length and bond angle
a=b=c
measurements. The interatomic lengths are listed in Table 2.
According to Table 2, an expansion of the tetrahedral site from the
CO

CFO- 8.375(4) 587.43(50) 17.7 23.0 15.0 2.33 4.6


0.4 CFO-0.4 to CFO-1.0 sample is observed. The Fe(1)–O bond length in-
CFO- 8.347(5) 581.56(60) 12.0 15.5 12.0 1.66 5.7
0.8
Table 2
CFO- 8.364(3) 585.12(40) 13.1 17.5 14.9 1.38 8.8
1.0 Interatomic lengths of Fe–O and Co–O bonds of cobalt ferrite powders in
CFO- 8.341(6) 580.30(70) 13.9 17.8 15.1 1.38 3.9 Fe(1)O4 tetrahedron and Fe(2)O6 octahedron.
1.2 Bond Site CFO-0.4 CFO-0.8 CFO-1.0 CFO-1.2

Therefore, the cluster formation by smaller crystallite size of CFO-1.2, Length (Å)

controlled by Co2+ diffusion, favors the formation of large particles Fe(1)–O (x4) Tetrahedron 1.8742 1.8954 1.9137 1.8203
(102.3 nm) compared to CFO-1.0 (74.5 nm). Fe(2)–O (x6) Octahedron 2.0592 2.0371 2.0334 2.0769
The lattice parameters and unit cell volume of cobalt ferrite pow- Valence (v.u.)
ders obtained via structural refinement are listed in Table 1. Fe3+–O Tetrahedron 2.93 2.77 2.63 3.39
The lattice parameter a varies at ca. 0.033 Å as a function of cobalt- Fe3+–O Octahedron 2.67 2.83 2.86 2.54
doping and it can be correlated with a decrease in the unit cell volume. Co2+–O Tetrahedron 2.44 2.31 2.20 2.83

A reduction in the unit cell volume indicates the formation of a substi- Co2+–O Octahedron 2.22 2.36 2.38 2.12

4
M.C.M. Ferreira et al. Optical Materials xxx (xxxx) 113043

creases at about 0.0395 Å, while the octahedral site in the CFO-0.4 to hedral site. The results are listed in Table 3. The bond angle variance for
CFO-1.0 samples contracts at ca. 0.0258 Å for the Fe(2)–O bond length. the octahedral site increases linearly from CFO-0.4 to CFO-1.0 at about
On the other hand, is worth noting that the CFO-1.2 sample presents a 8.32°, while it decreases at about 12.33° for CFO-1.2 compared to CFO-
contraction of the tetrahedral site at about 0.0934 Å and an expansion 1.0. On the other hand, the tetrahedral site is still unchanged in relation
of the octahedral site at around 0.0435 Å compared with the CFO-1.0 to Co2+-doping. According to Boxall et al. [36], the photoactivity in
sample. The preference of occupation of the Co2+ was analyzed in magnetite (Fe3O4) is due to the efficient electron exchange between
terms of bond valence. According to Brown [35], a bond valence theory Fe2+ and Fe3+ in the octahedral sites.
can be described in terms of the following equation (7): In addition, electronic transitions from ions tetrahedrally and octa-
hedrally coordinated, as in magnetite and spinel cobalt ferrites, gener-
(7) ate a set of intermediate gap states, which can extend its light absorp-

F
tion. Since, both conduction and valence bands are formed from the 3 d
where s is the bond valence, r0 and r are the standard and experi- state of Fe and the 2p state of oxygen, a shifting of the electrons to the
mental bond length, respectively, and b is an empirical parameter with 3 d state of Fe reduces the crystal field sufficiently and gives the forma-

OO
a value equal to 0.37. The r0 value reported by Brown [35] for Fe3+–O tion of extended 3 d gap states [39]. Therefore, in this work, the major
and Co2+–O standards were 1.759 Å and 1.692 Å, respectively. In addi- contribution to the photoactivity of cobalt ferrite can be assigned to the
tion, the oxidation state (Vi) in each crystallographic site of the spinel octahedron site distortion and Co2+ distribution between spinel atomic
structure was obtained by the summation of the bond valence (sij) ac- sites.
cording to equation (8) [35].
3.3. Chemical bonds and optical analysis
(8)

PR
The FTIR spectra of the cobalt ferrite powders are shown in Fig. 3a
The valence at the octahedron and tetrahedron sites of the spinel and FTIR and UV–vis spectroscopy results are listed in Table 4.
structure is listed in Table 2. The tetrahedral site assumes a valence All the samples showed similar bands at 3370 cm−1 and 1584 cm−1
close to +2.0 as Co2+-doped increases up to CFO-1.0. On the other assigned to O–H stretching (νOH) and O–H deforming (δOH), respectively
hand, the octahedron site assumes a valence of ca. +3.0 as cobalt- [40,41]. These bands indicate water molecules chemically adsorbed, in
doped increases in the same way. Thus, Fe3+ prefers octahedral coordi- the powders. A third band observed at around 1338 cm−1 is assigned to
nation, while Co2+ should occupies the tetrahedral coordination. These nitrate vibration [41], which can be derived from precursor reagents
results are in agreement with Sharma and coworkers [34]. However, adsorbed in the powders. The region between 1000 cm−1 and 250 cm−1
D
the difference in the valence of the octahedron site from the formal corresponds to the metal-oxygen (Me–O) bond. A band at around
charge +3.0 indicates that the preference of occupancy is not restricted 961 cm−1 is attributed to the Fe-Co alloy system [41]. Another band at
only to Fe3+, but also some Co2+ have occupancy at the octahedron site around 554 cm−1 is assigned to Co–O stretching vibration mode from
the octahedral site [40–42]. The last band at 350 cm−1 is related to the
TE

[36]. Therefore, there is some degree of inversion in the spinel struc-


ture. Above CFO-1.0, the spinel structure is preferentially inversed with Fe–O stretching vibration mode associated with the tetrahedral site
the octahedral and tetrahedral sites having +2.12 and + 3.39 valences [40,42]. These bands indicate the formation of the spinel CoFe2O4.
for the Co2+ and Fe3+ occupancy, respectively. These results indicate The optical absorption spectra from diffuse reflectance measure-
that in the range of CFO-0.4 – CFO-1.0 a normal spinel structure is fa- ments for the cobalt ferrite powders are shown in Fig. 3b. A broad band
vored with slight inversion, while above CFO-1.0 the spinel structure of cobalt ferrite powders is shown in the range of 220–800 nm, in the
EC

shows a major degree of inversion [15,37]. Based on the variance angle UV–Visible spectral region. This band is assigned to the overlapping 4A2
of the octahedral and tetrahedral sites [38], the distortion in both → 4T1 (4P) transition of the tetrahedral Co2+ and the Co2+ + Fe3+ →
atomic sites in the spinel structure of cobalt ferrite powders was mea- Co3+ + Fe2+ metal–metal charge transfer transition (MMCT) [43].
sured. The bond angles are listed in Table 3. Also, it can be seen from the CFO-1.0 and CFO-1.2 bands in the region
of 700–1300 nm (inset of Fig. 3b), related to the 4A2 → 4T1 (4F) crystal
field transition of the Co2+ ions in tetrahedral symmetry [43]. The ap-
(9) pearance of this band is in agreement with the results obtained from the
RR

structural refinement (section 3.2). Therefore, the lack of a vibrational


The variance angle (σ2) depends on the number of bonds (n), ith band in the region of 700–1300 nm for the CFO-0.4 and CFO-0.8 pow-
bond angles (θi), and ideal bond angle (θ0) of the atomic site. The ideal ders can be correlated with a spinel structure with some degree of in-
bond angle assumes 90° for the octahedral site and 109.47° for the tetra- version. For CFO-1.0, the Co2+ concentration is sufficient to create an
intense band in the region of 700–1300 nm, suggesting that the normal
spinel structure is favored. CFO-1.2 powder also presents an intense vi-
Table 3
CO

brational band in the region of 700–1300 nm, suggesting a normal


Bond angles (θ) of Fe(1)O4 tetrahedron and Fe(2)O6 octahedron of cobalt fer-
spinel structure with cobalt occupying the tetrahedral site. However,
rites.
according to rule of the valence sum for CFO-1.2, see section 3.2 discus-
Bond CFO-0.4 CFO-0.8 CFO-1.0 CFO-1.2
sion, Co2+ has a preference to occupy octahedral site with a valence of
θ (°) +2.12. In this sense, part of the Co2+ may migrate from tetrahedral to
the octahedral site in CFO-1.2, and the normal spinel structure becomes
O-Fe(1)-O (x6) 109.5 109.5 109.5 109.5 small inversion degree. Thus, these results can be correlate with a
O-Fe(2)-O (x6) 91.97 92.89 93.40 90.46
mixed spinel structure, i.e., CFO-1.2 can present both normal and in-
O-Fe(2)-O (x6) 88.03 87.10 86.61 89.54
O-Fe(2)-O (x3) 180.0 180.0 180.0 180.0 verted spinel with Co2+ occupying both tetrahedral and octahedral
Fe(1)-O-Fe(2) (x3) 123.9 123.2 122.9 124.9 sites.
Fe(2)-O-Fe(2) (x3) 91.94 92.83 93.30 90.46 Based on a direct transition (α = 1/2) [42,43], the energy gap of
σ2(oct) (°) 4.251 9.168 12.57 0.239 the cobalt ferrite powders was obtained from the Tauc plot method
σ2(tet) (.10−6 °) 1.200 1.200 1.200 1.200 (Fig. 3c). Considering values of energy gap derived from the best linear
range (2.25–3.25 eV), the intercept of the linear regression of the Tauc
The distortion measurement was calculated from equation (9) considering the plot with the energy axis determines the position of Eg Ref. [20] as
bond angles listed in Table 3.

5
M.C.M. Ferreira et al. Optical Materials xxx (xxxx) 113043

F
OO
PR
Fig. 3. a Infrared spectra of cobalt ferrites. b DR/UV–vis spectra in terms of the Kubelka-Munk theory. c Tauc plot according to a direct transition. d Linear regression
of Tauc plot in the range (2.25–3.25 eV).
D
Table 4 parameter photocatalytic efficiency (ζ) of the cobalt ferrite powders for
Attribution and wavenumber assignments of the Fe–O and Co–O vibration the oxidation of Rhd6G in water can be represented by equation (10)
bands obtained from FTIR spectra, and optical properties of samples, includ-
TE

[44].
ing maxima absorption wavelength (λabs-photon) and optical energy band gap
(Eg) obtained from DR/UV–vis spectra.
(10)
Powder Wavenumber (cm−1) λabs-photon (nm) Eg (eV)

Fe–O Co–O where C0 is the initial concentration of Rhd6G and Ct is the real-time
concentration of Rhd6G under UV light irradiation. A representative
EC

CFO-0.4 352 552 633 1.956


CFO-0.8 292 527 728 1.701 graphic of the photocatalytic efficiency of cobalt ferrites is shown in
CFO-1.0 310 550 633 1.958 Fig. 4f. The photocatalytic efficiency of cobalt ferrites seems to be re-
CFO-1.2 336 541 656 1.890 lated to the Co2+-doping. At alkaline pH, the CFO-0.8 powder has a
higher efficiency equal to 92.2%. However, the CFO-1.0 powder ex-
shown in Fig. 3d. The set of band gap values is listed in Table 4. The en- hibits an efficiency of 74.5% in acidic pH, 88.9% in neutral pH, and
ergy gap of cobalt ferrite synthesized by the Co-precipitation method 90.9% in alkaline pH. The increase in photoactivity under acidic condi-
RR

agrees with values of the literature [42,43]. The lower energy gap of tions can be associated with the increase of Co2+-doping: 65.7%,
cobalt ferrite powders indicates that the Co2+-doping gives rise a new 68.8%, 74.5%, and 76.2% for the CFO-0.4, CFO-0.8, CFO-1.0, and CFO-
level to the conduction band of ferrite and the electrons can be pro- 1.2, respectively. On the other hand, the efficiency in the Co2+-doping
moted more easily from the valence band to these ions' levels, favoring higher than CFO-1.0 decreases in neutral and alkaline pH. The kinetics
the cobalt ferrite powders’ response in the visible region. of Rhd6G degradation reaction over cobalt ferrite catalysts was investi-
gated in terms of the Langmuir-Hinshelwood mechanism [1]. In this
CO

3.4. Photocatalysis mechanism, the rate of reaction (r) is proportional to the surface loaded
with pollutants (θ) according to equation (11).
The photocatalytic activity of cobalt ferrites was evaluated by the
(11)
photodegradation of Rhd6G as a pollutant model. According to diffuse
reflectance data, cobalt ferrite has a large absorption band in the
where kr is the reaction-rate constant, Kads the adsorption constant
UV–Vis region. The photodegradation of Rhd6G (1 L, 10.4 μmol L−1)
according to the Langmuir model, and Ceq is the concentration of the
under irradiation and in the absence of catalyst is shown in Fig. 4a. The
pollutant in the equilibrium conditions. As the Ceq value is lower
Rhd6G has a minimum degradation in the absence of catalyst, being
(≤10−3 mol L−1) for diluted systems, a simplification is made consider-
more susceptible to degradation in neutral (pH 7.0) to alkaline (pH 9.0)
ing . Thus, equation (11) is simplified to first-order kinet-
conditions. Fig. 4b–e shows the adsorption of the series of Rhd6G solu-
ics described by equation (12).
tions in the presence of the cobalt ferrite powders (100 mg) under light
irradiation for 240 min. (12)
In the presence of cobalt ferrites, Rhd6G concentration decreases
gradually with the irradiation time. The kinetics of degradation be-
comes more pronounced when the reaction occurs at alkaline pH. The

6
M.C.M. Ferreira et al. Optical Materials xxx (xxxx) 113043

F
OO
PR
D
TE
EC
RR

Fig. 4. Kinetics of Rh6G degradation under UV-irradiation: a Direct photolysis, b CFO-0.4, c CFO-0.8, d CFO-1.0, e CFO-1.2. f Photodegradation efficiency of cobalt
ferrite powders.
CO

where kapp is the first-order apparent rate constant which can be ob- rate constant (kapp), the cobalt plays a crucial role in the photoactivity
tained from the linear regression of the kinetic data of degradation, as a of cobalt ferrite powders. In a broad sense, the increase in the degrada-
function of the irradiation time, according to equation (13). tion can be derived from aspects such as particle size, structure, and re-
active sites [1], as well as surface characteristics.
(13) The higher photoactivity level of cobalt ferrite powders can be due
to the intrinsic characteristics and surface area of nanoparticle clusters.
where Ceq is the initial concentration after achieving the equilibrium In this way, cobalt in tetrahedral coordination, as pointed out by the va-
of adsorption in the absence of light, and Ct is the concentration at the lence shown in Table 2, favors the photoactivity of CFO-0.4, CFO-0.8,
time t of irradiation. and CFO-1.0 powders in neutral and alkaline media. However, the octa-
Fig. 5 shows the linear regressions of the kinetic data from Fig. 4b–e, hedral coordination cobalt, as in CFO-1.2, benefits the photo activity in
and a set of the kinetic parameters obtained for the Rhd6G photodegra- the acidic medium. In this context, the higher CFO-0.8 photoactivity
dation is listed in Table 5. over CFO-1.0 can be understood by means of cobalt occupation in the
According to Fig. 5, Rhd6G photodegradation follows a pseudo-first- tetrahedral site and its control of the Rhd6 adsorption over different pH
order reaction with good linear fitting. From the first-order apparent

7
M.C.M. Ferreira et al. Optical Materials xxx (xxxx) 113043

F
OO
PR
Fig. 5. Linear regression from the kinetic data in Fig. 4: a CFO-0.4. b CFO-0.8. c CFO-1.0. d CFO-1.2.
D
Table 5 both tetrahedral and octahedral sites. For this reason, as shown in Fig.
List of kinetics parameters of Rhd6G photodegradation, in distinct pH values. 3b for CFO-1.0 and CFO-1.2, the octahedral site loses its distortion and
becomes more symmetrical with angles close to 90°.
TE

Powder pH kapp (min−1) x10−3 R2

CFO-0.4 Acidic 3.99 0.9562 4. Conclusions


Neutral 6.04 0.9705
Alkaline 9.49 0.9977 Cobalt ferrite nanoparticles with stoichiometry Co0.4Fe2.6O4,
CFO-0.8 Acidic 4.76 0.9832 Co0.8Fe2.2O4, CoFe2O4, and Co1.2Fe1.8O4 have been synthesized by the
Neutral 6.87 0.9313
wet chemical route, a Co-precipitation method, at 80 °C. The nanoparti-
EC

Alkaline 11.1 0.9931


cles show spherical morphology arranged in cluster form. Structural
CFO-1.0 Acidic 6.15 0.9759
Neutral 9.07 0.9977 analysis from XRD confirmed the synthesis of a cobalt ferrite single-
Alkaline 10.23 0.9991 phase with a spinel structure of cubic symmetry . Structural re-
CFO-1.2 Acidic 5.65 0.9661 finement by the Rietveld method showed an inverse spinel structure for
Neutral 6.93 0.9702 CFO-0.4, while for CFO-0.8 and CFO-1.0 a normal structure was more
Alkaline 9.30 0.9940
favorable. However, for CFO-1.2, the inverse spinel structure is again
RR

predominant. Both FTIR and UV–visible spectroscopies showed that the


conditions. Accordingly, the photoactivity of cobalt ferrites has an inti- Co2+ is tetrahedral coordination in the spinel structure for CFO-1.0 and
mate relationship with the cobalt tetrahedrally coordinated. CFO-1.2 samples. The CFO-1.0 sample showed the best photoactivity
As highlighted in the structural characterization, the octahedral site correlation in the degradation of the RhD6G in acidic (74.5%), neutral
becomes more distorted with cobalt concentration increase in this site (88.9%), and alkaline media (90.9%). The electronic excitation and mi-
and makes Fe3+ migrate to the tetrahedral site, in according to Table gration on the surface of nanoparticles have an intrinsic correlation to
data and discussion. From this point, the reaction rate increases at a octahedral site distortion and Co2+ occupation at both octahedron and
CO

neutral pH from 6.04 × 10−3 min−1 (CFO-0.4) to 9.07 × 10−3 min−1 tetrahedron sites. According to the results, superficial reactions in an al-
(CFO-1.0); and at alkaline pH from 9.49 × 10−3 min−1 (CFO-0.4) to kaline medium are favorably catalyzed by cobalt ferrite with a predom-
11.1 × 10−3 min−1 (CFO-0.8), and 10.23 × 10−3 min−1 (CFO-1.0). This inant normal spinel structure. On the other hand, reactions under acidic
behavior indicates a narrow correlation with the octahedral site activity conditions are best catalyzed with cobalt ferrites of inverse spinel type
as it becomes more distorted (see Table 3). Earlier, similar effects in fer- structure. These finds could be beneficial for designing new ferrite cata-
roelectric perovskites [22] and tetragonal tungsten bronze structures lysts for different waste water environment treatments.
[45] have been reported. Formation of reactive species can be derived
from apical oxygen atoms in octahedral sites, which is not as efficient in CRediT authorship contribution statement
the acid medium as in the alkaline medium [46], and thus, a decrease in
photoactivity is expected. On the other hand, the tetrahedral site con- M.C.M. Ferreira : Investigation. F.R. Praxedes : Investiga-
trols the photoactivity of cobalt ferrites in the acidic medium. tion. M.A.L. Nobre : Conceptualization, Investigation. S. Lanfredi :
The rate constant of CFO-1.2 shows a slight change in the degrada- Conceptualization, Investigation.
tion reaction of Rhd6G in the acidic medium compared to the neutral
and alkaline medium. The structural and optical analysis of CFO-1.2
shows that Co2+ at higher concentrations has a preference to occupy

8
M.C.M. Ferreira et al. Optical Materials xxx (xxxx) 113043

Declaration of competing interest https://doi.org/10.1016/0921-4526(93)90108-I.


[19] J. Rodríguez-Carvajal, An Introduction to the Program FullProff 2000
(Laboratorie Léon Brillouin, CEACNRS, Saclay), 2001 France.
The authors declare that they have no known competing financial [20] K. Momma, F. Izumi, VESTA 3 for three-dimensional visualization of crystal,
interests or personal relationships that could have appeared to influ- volumetric and morphology data, J. Appl. Crystallogr. 44 (2011) 1272–1276,
https://doi.org/10.1107/S0021889811038970.
ence the work reported in this paper.
[21] P. Makuła, M. Pacia, W. Macyk, How to correctly determine the band gap energy
of modified semiconductor photocatalysts based on UV-Vis spectra, J. Phys. Chem.
Data availability Lett. 9 (2018) 6814–6817, https://doi.org/10.1021/acs.jpclett.8b02892.
[22] F.R. Praxedes, M.A.L. Nobre, P.S. Poon, J. Matos, S. Lanfredi, Nanostructured
KxNa1-xNbO3 hollow spheres as potential materials for the photocatalytic treatment
No data was used for the research described in the article. of polluted water, Appl. Catal. B Environ. 298 (2021) 120502, https://doi.org/
10.1016/j.apcatb.2021.120502.

F
Acknowledgment [23] R.B. McClurg, R.C. Flagan, Critical comparison of droplet models in
homogeneous nucleation theory, J. Colloid Interface Sci. 201 (2) (1998) 194–199,
https://doi.org/10.1006/jcis.1997.5379.
F.R. Praxedes thanks “Coordenação de Aperfeiçoamento de Pessoal [24] B.J. McCoy, Distribution kinetics modeling of nucleation, growth, and

OO
de Nível Superior (CAPES)” - Finance Code 001. S. Lanfredi acknowl- aggregation processes, Ind. Eng. Chem. Res. 40 (2001) 5147–5154, https://
doi.org/10.1021/ie001034i.
edges the Brazilian projects: Grant# 2002/05997-9, 2007/03510-9, [25] M. Sedlacik, V. Pavlinek, P. Peer, P. Filip, Tailoring the magnetic properties and
2014/11189-0, 2019/06623-6, 2020/00781-6 from São Paulo Re- magnetorheological behavior of spinel nanocrystalline cobalt ferrite by varying
search Foundation (FAPESP) and CNPq for financial support. V.P. Car- annealing temperature, Dalton Trans. 43 (2014) 6919–6924, https://doi.org/
10.1039/c4dt00166d.
valho-Jr for DR/UV–vis measurements.
[26] B.G. Toksha, S.E. Shirsath, M.L. Mane, S.M. Patange, S.S. Jadhav, K.M. Jadhav,
Autocombustion high-temperature synthesis, structural, and magnetic properties of
References CoCrxFe2-xO4 (0 ≤ x ≤ 1.0), J. Phys. Chem. C 115 (43) (2011) 20905–20912,

PR
https://doi.org/10.1021/jp205572m.
[1] J.M. Herrmann, Photocatalysis fundamentals revisited to avoid several [27] M. El-Hilo, A.A. Dakhel, Structural and magnetic properties of Mn-doped ZnO
misconceptions, Appl. Catal. B Environ. 99 (2010) 461–468, https://doi.org/ powders, J. Magn. Magn Mater. 323 (2011) 2202–2205, https://doi.org/10.1016/
10.1016/j.apcatb.2010.05.012. j.jmmm.2011.03.031.
[2] Y. Zhang, Z.R. Tang, X. Fu, Y.J. Xu, TiO2 Graphene nanocomposites for gas-phase [28] M. Casas-Cabanas, M.R. Palacín, J. Rodríguez-Carvajal, Microstructural analysis
photocatalytic degradation of volatile aromatic pollutant: is TiO2-graphene truly of nickel hydroxide: anisotropic size versus stacking faults, Powder Diffr. 20 (2005)
different from other TiO2-carbon composite materials, ACS Nano 4 (12) (2014) 334–344, https://doi.org/10.1154/1.2137340.
7303–7314, https://doi.org/10.1021/nn1024219. [29] D. Karthickraja, S. Karthi, G.A. Kumar, D.K. Sardar, G.C. Dannangoda, K.S.
[3] J. Li, C. Yu W. Fang, W. Zhou, L. Zhu, Y. Xie, Ag-based semiconductor Martirosyan, E.K. Girija, Fabrication of core-shell CoFe2O4@HAp nanoparticles: a
photocatalysts in environmental purification, Appl. Surf. Sci. 358 (A) (2015) novel magnetic platform for biomedical applications, New J. Chem. 43 (2019)
D
46–56, https://doi.org/10.1016/j.apsusc.2015.07.139. 13584–13593, https://doi.org/10.1039/c9nj02510c.
[4] P. Dong, G. Hou, X. Xi, R. Shao, F. Dong, WO3-based photocatalysts: morphology [30] K. Maaz, A. Mumtaz, S.K. Hasanain, A. Ceylan, Synthesis and magnetic
control, activity enhancement and multifunctional applications, Environ. Sci. Nano properties of cobalt ferrite (CoFe2O4) nanoparticles prepared by wet chemical
4 (3) (2017) 539–557, https://doi.org/10.1039/c6en00478d. route, J. Magn. Magn Mater. 308 (2007) 289–295, https://doi.org/10.1016/
[5] Y. Liu, Y. Song, Y. You, X. Fu, J. Wen, X. Zheng, NiFe2O4/g-C3N4 heterojunction j.jmmm.2006.06.003.
TE

composite with enhanced visible-light photocatalytic activity, J. Saudi Chem. Soc. [31] Y. Zhang, Z. Yang, D. Yin, Y. Liu, C. Fei, R. Xiong, J. Shi, G. Yan, Composition and
22 (4) (2018) 439–448, https://doi.org/10.1016/j.jscs.2017.08.002. magnetic properties of cobalt ferrite nano-particles prepared by the co-
[6] Y. Lu, B. Ren, S. Chang, W. Mi, J. He, W. Wang, Achieving effective control of the precipitation method, J. Magn. Magn Mater. 322 (2010) 3470–3475, https://
photocatalytic performance for CoFe2O4/MoS2 heterojunction via exerting external doi.org/10.1016/j.jmmm.2010.06.047.
magnetic fields, Mater. Lett. 260 (2020) 126979, https://doi.org/10.1016/ [32] C.L. Muhich, V.J. Aston, R.M. Trottier, A.W. Weimer, C.B. Musgrave, First-
j.matlet.2019.126979. Principles analysis of cation diffusion in mixed metal ferrite spinels, Chem. Mater.
[7] A. Sudhaik, P. Raizada, P. Shandilya, P. Singh, Magnetically recoverable 28 (2016) 214–226, https://doi.org/10.1021/acs.chemmater.5b03911.
graphitic carbon nitride and NiFe2O4 based magnetic photocatalyst for degradation [33] S.K. Gore, S.S. Jadhav, V.V. Jadhav, S.M. Patange, M. Naushad, R.S. Mane, K.H.
EC

of oxytetracycline antibiotic in simulated wastewater under solar light, J. Environ. Kim, The structural and magnetic properties of dual phase cobalt ferrite, Sci. Rep. 7
Chem. Eng. 6 (2018) 3874–3883, https://doi.org/10.1016/j.jece.2018.05.039. (1) (2017) 2524, https://doi.org/10.1038/s41598-017-02784-z.
[8] K.K. Kefeni, B.B. Mamba, T.A. Msagati, Application of spinel ferrite nanoparticles [34] R. Sharma, P. Thakur, M. Kumar, N. Thakur, N.S. Negi, P. Sharma, V. Sharma,
in water and wastewater treatment: a review, Separ. Purif. Technol. 188 (2017) Improvement in magnetic behaviour of cobalt doped magnesium zinc nano-ferrites
399–422, https://doi.org/10.1016/j.seppur.2017.07.015. via co-precipitation route, J. Alloys Compd. 684 (2016) 569–581, https://doi.org/
[9] A. Al-Anazi, W.H. Abdelraheem, K. Scheckel, M.N. Nadagouda, K. O’Shea, D.D. 10.1016/j.jallcom.2016.05.200.
Dionysiou, Novel franklinite-like synthetic zinc-ferrite redox nanomaterial: [35] I.D. Brown, Recent developments in the methods and applications of the bond
synthesis, and evaluation for degradation of diclofenac in water, Appl. Catal. B valence model, Chem. Rev. 109 (2009) 6858–6919, https://doi.org/10.1021/
RR

Environ. 275 (2020) 119098, https://doi.org/10.1016/j.apcatb.2020.119098. cr900053k.


[10] E. Ferdosi, H. Bahiraei, D. Ghanbari, Investigation the photocatalytic activity of [36] C. Boxall, G. Kelsall, Z. Zhang, Photoelectrophoresis of colloidal iron oxides. Part
CoFe2O4/ZnO and CoFe2O4/ZnO/Ag nanocomposites for purification of dye 2.—magnetite (Fe3O4), J. Chem. Soc., Faraday Trans. 2 (1998) 43527–43780,
pollutants, Separ. Purif. Technol. 211 (2019) 35–39, https://doi.org/10.1016/ https://doi.org/10.1039/FT9969200791.
j.seppur.2018.09.054. [37] Y.H. Hou, Y.J. Zhao, Z.W. Liu, H.Y. Yu, X.C. Zhong, W.Q. Qiu, D.C. Zeng, L.S.
[11] Z. Jia, D. Ren, Y. Liang, R. Zhu, A new strategy for the preparation of porous zinc Wen, Structural, electronic and magnetic properties of partially inverse spinel
ferrite nanorods with subsequently light-driven photocatalytic activity, Mater. Lett. CoFe2O4: a first-principles study, J. Phys. D Appl. Phys. 43 (2010) 445003, https://
65 (2011) 3116–3119, https://doi.org/10.1016/j.matlet.2011.06.101. doi.org/10.1088/0022-3727/43/44/445003.
[38] K. Robinson, G.V. Gibbs, P.H. Ribbe, Quadratic elongation: a quantitative
CO

[12] J.K. Rajput, G. Kaur, Synthesis and applications of CoFe2O4 nanoparticles for
multicomponent reactions, Catal. Sci. Technol. 4 (2014) 142–151, https://doi.org/ measure of distortion in coordination polyhedra, Science 172 (1971) 567–570,
10.1039/c3cy00594a. https://doi.org/10.1126/science.172.3983.567.
[13] G. El-Shobaky, A. Turky, N. Mostafa, S. Mohamed, Effect of preparation [39] P. Mishra, S. Patnaik, K. Parida, An overview of recent progress on noble metal
conditions on physicochemical, surface and catalytic properties of cobalt ferrite modified magnetic Fe3O4 for photocatalytic pollutant degradation and H2
prepared by coprecipitation, J. Alloys Compd. 493 (2010) 415–422, https:// evolution, Catal. Sci. Technol. 9 (2019) 916–941, https://doi.org/10.1039/
doi.org/10.1016/j.jallcom.2009.12.115. c8cy02462f.
[14] V. Pillai, D. Shah, Synthesis of high-coercivity cobalt ferrite particles using water- [40] J.B. Silva, W. De Brito, N.D.S. Mohallem, Influence of heat treatment on cobalt
in-oil microemulsions, J. Magn. Magn Mater. 163 (1996) 243–248, https://doi.org/ ferrite ceramic powders, Mater. Sci. Eng B-Solid. 112 (2004) 182–187, https://
10.1016/S0304-8853(96)00280-6. doi.org/10.1016/j.mseb.2004.05.029.
[15] K.E. Sickafus, J.M. Wills, N.W. Grimes, Structure of spinel, J. Am. Ceram. Soc. 82 [41] D. Karthickraja, S. Karthi, G.A. Kumar, D.K. Sardar, G.C. Dannangoda, K.S.
(1999) 3279–3292, https://doi.org/10.1111/j.1151-2916.1999.tb02241.x. Martirosyan, E.K. Girija, Fabrication of core-shell CoFe2O4@HAp nanoparticles: a
[16] I. Safarik, K. Horska, K. Pospiskova, Z. Maderova, M. Safarikova, Microwave novel magnetic platform for biomedical applications, New J. Chem. 43 (2019)
assisted synthesis of magnetically responsive composite materials, IEEE Trans. 13584–13593, https://doi.org/10.1039/c9nj02510c.
Magn. 49 (2012) 213–218, https://doi.org/10.1109/TMAG.2012.2221686. [42] M.H. Habibi, H.J. Parhizkar, FTIR and UV-vis diffuse reflectance spectroscopy
[17] D. Zhao, X. Wu, H. Guan, E. Han, Study on supercritical hydrothermal synthesis studies of the wet chemical (WC) route synthesized nano-structure CoFe2O4 from
of CoFe2O4 nanoparticles, J. Supercrit. Fluids 42 (2) (2007) 226–233, https:// CoCl2 and FeCl3, Spectrochim. Acta Mol. Biomol. Spectrosc. 127 (2014) 102–106,
doi.org/10.1016/j.supflu.2007.03.004. https://doi.org/10.1016/j.saa.2014.02.090.
[18] J. Rodríguez-Carvajal, Recent advances in magnetic structure determination by [43] I.H. Gul, A. Maqsood, M. Naeem, M.N. Ashiq, Optical, magnetic and electrical
neutron powder diffraction, Phys. B: Phys. Condens. Matter 192 (1993) 55–69, investigation of cobalt ferrite nanoparticles synthesized by co-precipitation route,
J. Alloys Compd. 507 (2010) 201–206, https://doi.org/10.1016/

9
M.C.M. Ferreira et al. Optical Materials xxx (xxxx) 113043

j.jallcom.2010.07.155. [46] J. Matos, J. Arcibar-Orozco, P.S. Poon, G. Pecchi, J.R. Rangel-Mendez, Influence
[44] F. Wang, L.X. Song, Y. Teng, J. Xia, Z.Y. Xu, W.P. Wang, Synthesis, structure, of phosphorous upon the formation of DMPO-•OH and POBN-O2•− spin-trapping
magnetism and photocatalysis of α-Fe2O3 nanosnowflakes, RSC Adv. 9 (2019) adducts in carbon-supported P-promoted Fe-based photocatalysts, J. Photochem.
35372–35383, https://doi.org/10.1039/c9ra07490b. Photobiol., A: Chem 391 (2020) 112362, https://doi.org/10.1016/
[45] J. Matos, S. Lanfredi, R. Montaña, M.A.L. Nobre, M.C. Fernández de Córdoba, j.jphotochem.2020.112362.
C.O. Ania, Photochemical reactivity of apical oxygen in KSr2Nb5O15 materials for
environmental remediation under UV irradiation, J. Colloid Interface Sci. 496
(2017) 211–221, https://doi.org/10.1016/j.jcis.2017.02.028.

F
OO
PR
D
TE
EC
RR
CO

10

View publication stats

You might also like