Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Catalysis Communications 174 (2023) 106567

Contents lists available at ScienceDirect

Catalysis Communications
journal homepage: www.elsevier.com/locate/catcom

Low power photo-assisted catalytic degradation of azo dyes using 1-D BiOI:
Optimization of the key physicochemical features
Marzieh Nourzad a, Aliakbar Dehghan a, Zohreh Niazi b, Dimitrios A. Giannakoudakis c, *,
Mojtaba Afsharnia d, Mariusz Barczak e, Ioannis Anastopoulos f, Konstantinos S. Triantafyllidis c,
Mahmoud Shams a, *
a
Social Determinants of Health Research Center, Mashhad University of Medical Sciences, Mashhad, Iran
b
Chemistry Department, Faculty of Science, Ferdowsi University of Mashhad, Mashhad, Iran
c
Department of Chemistry, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
d
Department of Environmental Health Engineering, Faculty of Health, Social Development and Health Promotion Research Center, Gonabad University of Medical
Sciences, Gonabad, Iran
e
Institute of Chemical Sciences, Faculty of Chemistry, Maria Curie-Sklodowska University, Maria Curie-Sklodowska Sq. 3, 20-031 Lublin, Poland
f
Department of Agriculture, University of Ioannina, UoI Kostakii Campus, 47100 Arta, Greece

A R T I C L E I N F O A B S T R A C T

Keywords: Flower-like spherical 1-D-BiOI nanoparticles were solvothermally synthesized and evaluated for the photo­
Photocatalytic water treatment catalytic degradation of the azo dyes Methyl Orange (MO), Methylene Blue (MB), and Eriochrome Black-T (EBT)
Low power LED using ultra-low-power LED light. The soft white-light bulb positively affected the photo-assisted remediation
Azo dye
efficiency. The tests showed that 1-D-BiOI removal efficiency against 12 mg/L dyes MO, MB, and EBT were 47.4,
1-D BiOI
66.1, and 86.7%, respectively. The optimum experimental parameters were determined by modeling and opti­
Bismuth oxy‑iodine nanomaterials
mization using RSM. Scavenging tests indicated that •OH are the main responsible species. COD was increased
initially during light irradiation and then decreased by dye mineralization.

1. Introduction the stability of the azo group (-N=N-). Methyl Orange (MO), Methylene
Blue (MB), and Eriochrome Black T (EBT) are among the persistent,
In recent years, with the rapid progress of industrial, urban, and toxic, and recalcitrant azo dyes [10,11].
agricultural activities, a large number of pollutants have been released MO is an anionic dye or sulfonated azo dye, used not only in the
into water resources [1]. Of pollutants, organics have attracted much textile dyeing industry but also as a pH indicator in biomedical. Azo dye
attention because they pose a serious threat to human health and the metabolites including MO were reported to be carcinogenic, teratogenic,
environment. Organic pollutants are discharged from a wide spectrum of mutagenic, and highly recalcitrant. Thus, MO-containing wastewater
industries i.e. leather and textile fabrics, paper, cosmetics, and paint should be decolorized and detoxified before being discharged into the
[2–4]. Dyes are significant contaminants that are used widely in the environment [12,13].
textile, pharmaceutical, food, cosmetics, plastics, paint, ink, photo­ MB is a cationic dye thiazine-dye with pervasive applications as
graphic, and paper industries [5]. Dyes can classify based on their colorants, treating fish’s fungal infection; inhibiting nitrification as well
structural properties as acidic, reactive, basic, disperse, azo, diazo, as medicinal aids. The marine pollution of MB is reported to exert
anthraquinone-based, and metal-complex dyes [6]. The heterocyclic abnormal metabolism, morphological deterioration, and other dys­
nature and the presence of carcinogen benzidine and naphthalene in the functions. The unregulated dumping of these dyes and auxiliary chem­
structure of most dyes bring a challenge as they are introduced into the icals into water bodies triggers a dreadful impact on marine ecology.
environment [7,8]. Moreover, the presence of these dyes in water can Adverse effects of acute exposure to MB result in health ailments like
cause a significant reduction in oxygen levels which in turn disturbs the necrosis, nausea, vomiting, eye irritation, and diarrhea. Above all, MB
marine life cycle [9]. dye is exerting
Of dyes, azo dyes are widely used due to their water solubility and recalcitrance effect to biodegrade. Therefore, MB dye removal from

* Corresponding authors.
E-mail address: dagchem@gmail.com (D.A. Giannakoudakis).

https://doi.org/10.1016/j.catcom.2022.106567
Received 13 September 2022; Received in revised form 8 November 2022; Accepted 19 November 2022
Available online 21 November 2022
1566-7367/© 2022 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
M. Nourzad et al. Catalysis Communications 174 (2023) 106567

wastewater streams is of vital concern [14]. was found to be comparatively higher than the pure BiOI and AgVO3. Liu
On the other hand, EBT is an anionic and soluble organic dye in et al. [43] synthesized BiOI/BiOCl heterojunctions with the assistance of
water [15]. Out of all major toxins, the higher influx of dyes especially reactive ionic liquid (IL) 1-propyl-3-methylimidazolium iodide for the
EBT from different applications such as textiles, plastic and leather in­ degradation of RhB. Photocatalytic performance evaluation results
dustry, medical industry, automobiles, and paper printing has resulted demonstrated that the photocatalytic activity of BiOI/BiOCl is several
in the deterioration of the environment. The extensive usage of EBT as times that of the bare BiOCl and BiOI.
an indicator in complexometric titrations and biological sample mark­ Jiang et al. [44] synthesized a series of BiOI/MIL− 125(Ti) compos­
ings has further enhanced its exposure rate. The excessive discharge of ites. The optimized BiOI/MIL− 125(Ti) composite exhibited excellent
EBT into water resources has affected the well-being of water organisms photocatalytic activity for the degradation of Tetracycline (TC). Guan
by affecting essential parameters such as the re‑oxygenation ability of et al. [45] synthesized a flower-like BiOI1-x with rich iodine vacancies,
water bodies and the photosynthetic abilities of the phytoplanktons and which exhibited photocatalytic activity for the removal of gaseous
aquatic plants. In addition, the excessive dosage of EBT has caused un­ mercury. Ji and coworkers [46] reported the fabrication of Bi4O5I2 ul­
desirable anomalies such as astigmatism and skin allergies. Therefore, it trathin hollow nanotubes with increased photocatalytic activity towards
is essential to remove dyes by effective and affordable techniques before BPA. Yang et al. [47] prepared a novel core/shell Bi12O15Cl6/BiOI
their discharge into the water bodies [16–20]. photocatalyst, with n-p junction heterojunction, by using p-type BiOI
Until now, various technologies including electrocatalysis, ozona­ nanosheet coating with lamellar n-type Bi12O15Cl6. The optimal sample
tion, ultrafiltration, chemical ion exchange, adsorption, and photo­ Bi12O15Cl6/BiOI degraded 90.3% MO in 90 min, 75.5% TC in 180 min,
catalysis, have been used to purify wastewater [21,22]. However, they and destroyed E. coli in 40 min under LED light illumination. Arumugam
mostly suffer from the generation of secondary pollutants, low removal et al. [48] synthesized BiOI by a hydro/solvothermal process using
efficiency at low concentrations, high operational costs, and limited different solvents for the reduction of Cr (VI) under visible light irradi­
commercial applications [23]. ation. Almost complete reduction of Cr(VI) was achieved within 60 min.
Advanced oxidation processes (AOPs) are recently introduced as Phuruangrat et al. [49] studied the effect of precursor solution pH on the
promising wastewater treatment techniques to remove various organic structural and photocatalytic performance of Bi2WO6. The study indi­
contaminants. AOPs decompose organic pollutants by generating cated that pH was a major variable in the synthesis and the highest RhB
oxidizing radicals such as OH•, O•- 2 , and HOO . Owing to the strong

degradation occurred when the catalyst was prepared at pH 2.
oxidizing potential, free radicals can mineralize most organic pollutants Bismuth-based photocatalysts have attracted widespread attention
to form benign species such as CO2 and H2O [24]. for wastewater treatment due to their low toxicity, premium stability,
Among AOPs, semiconductor photocatalysis is the most efficient special electronic structure, excellent optical response, remarkable
approach due to its simplicity, low cost, high degradation efficiency, and photo-oxidation ability, interlayer static electric field, and abundant
working under ambient conditions [25–30]. To be an effective photo­ structural defects [50–52]. Especially, bismuth oxyhalides (BiOX, X =
catalyst, the materials should have narrow bandgap energy, slow charge Cl, Br, I) have been widely studied as promising photocatalysts for water
recombination, fast charge transportation, and effective interfacial in­ treatment owing to their tunable band gap values, anisotropic structure,
teractions [31]. high stability, and excellent photocatalytic activity [53,54]. Interest­
Tremendous interest has been in the fabrication of bismuth-based ingly, BiOX represented a high level of photocatalytic activity under
photocatalysts in recent years [32–35]. ZnO/Bi2WO6 (ZBW) hetero­ ultraviolet or visible light illumination. Uniquely, bismuth oxy‑iodine
structures were prepared by Liu et al. to abate MB [36]. The obtained (BiOI), is a p-type semiconductor with a narrow bandgap (about 1.7–1.9
ZBW heterostructures exhibited superior photocatalytic activity than eV) that is superior among BiOX. BiOI absorbs light in a wide range of
that of pure ZnO and Bi2WO6 under ultraviolet light irradiation. Liu visible spectrums because of the special layered crystal structure [55,56]
et al. [37] prepared a novel double-Z-scheme Bi2S3/BiVO4/TiO2 (BVT) [57].
heterostructure for photocatalytic degradation of MO under visible light Solvothermal, hydrolysis, and electrospinning synthesis methods
irradiation. The results showed that optimized BVT possesses the best have been successfully used to prepare BiOI with controllable mor­
degradation efficiency compared to bare TiO2 nanorod. Chen et al. [38] phologies i.e. nanorods, nanoplates, and microflowers [58]. Jiang et al.
incorporated graphene oxide into the structure of bismuth halides and [59] synthesized a hollow flower-like BiOI photocatalyst (h-BiOI) via a
realized the photocatalytic activity enhanced significantly against facile solvothermal method. The h-BiOI displayed a complete degrada­
crystal violet and Rhodamine B (RhB). Interestingly, bismuth-based tion efficacy towards RhB under visible light irradiation. Dehghan et al.
heterojunctions also opted in research on the conversion of carbon di­ [60] synthesized two different BiOI by
oxide to methane. Lin et al. incorporated graphitic carbon nitride into hydrolysis (BiOI-H) and solvothermal (BiOI-ST) methods for photo­
bismuth compounds to fabricate BiOF/BiOI/Bi26O38F2/g-C3N4 and catalytic degradation of Tetracycline Hydrochloride (TCH). The BiOI-ST
Bi12O17Cl2/Bi5O7Br/Bi5O7I/g-C3N4. These catalysts exhibited superb exhibited higher photocatalytic activity in TCH degradation than BiOI-
activity for CO2 conversion and also CV degradation under visible light H. Lan et al. [61] fabricated BiOI flower-like spherical particles by a
irradiation [39,40]. hydrothermal method. Flower-like microsphere BiOI revealed higher
The novel green recyclable BiOX-ceramic fiber (CerF) photocatalysts photocatalytic activity towards 2,2-bis (4-hydroxyphenyl) propane
were synthesized via immobilization of Azadirachta indica leaf extract- degradation under visible light irradiation, compared to commercial
mediated BiOX nanoflowers on the Al2O3-based ceramic fiber sheet TiO2 P25.
[41]. The as-prepared recyclable BiOX-CerF samples i.e. BiOCl-CerF, In this work, we prepared a novel 1-D BiOI with premium photo­
BiOBr-CerF, and BiOI-CerF exhibited excellent photocatalytic perfor­ catalytic activity through a facile solvothermal method in N,N-dime­
mance and stability for MO, Bisphenol A (BPA), and Ampicillin degra­ thylformamide (DMF) medium. The photocatalytic activity of 1-D BiOI
dation under visible-light irradiation. Yuan et al. [42] prepared a series under an ultra-low powered LED source was evaluated by degrading azo
of ultrathin BiOX (X = Cl, Br) nanosheets (NSs) with surface oxygen dyes MO, MB, and EBT as model pollutants. Moreover, we developed a
vacancies through laser irradiation of bulk BiOX materials in deionized dye degradation model to predict the removal efficiency in the photo­
water without any organic additive. The engineered ultrathin BiOX NSs catalytic system. The model was then optimized for the highest degra­
exhibited excellent visible-light photocatalytic performances on the dation of the contaminant. Mineralization of the organic dye and the
degradation of RhB. Bavani et al. [24] fabricated a visible-light-driven active radical species determined by chemical oxygen demand (COD),
AgVO3/BiOI nanocomposite photocatalyst by using a facile hydrother­ and radical trapping tests, respectively.
mal method for the degradation of RhB under visible light irradiation.
The photocatalytic efficiency of optimized AgVO3/BiOI nanocomposite

2
M. Nourzad et al. Catalysis Communications 174 (2023) 106567

2. Experimental section 2.4. Dye photodegradation preliminary tests

2.1. Materials The photocatalytic activity of 1-D BiOI was evaluated by degrading
model pollutants (MO, MB, and EBT) under ultra-low powered LED
Bismuth nitrate hydrated (Bi(NO3)3⋅5H2O), potassium iodide (KI), sources. 1-D BiOI (400 mg/L) was added to solutions containing 12 and
DMF, and the three model pollutants/dyes MO, MB, and EBT were 15 mg/L dye and the mixtures were stirred in the dark for 30 min to
purchased from Sigma-Aldrich. Crucial details regarding the model achieve the adsorption-desorption equilibrium. Subsequently, the stir­
pollutants are shown in Table 1. AgNO3, ascorbic acid (AA), ethanol, and ring continued under irradiation by visible light for 60 min using 15-
isopropyl alcohol (IPA) were also purchased from Sigma-Aldrich. All the watt (W) LED bulbs. The dye removal after adsorption time, and dur­
reagents were used directly without further purification. Double ing photocatalysis degradation were monitored at regular time intervals.
distilled water (dDW) was used to prepare all the solutions/dispersions Dye concentrations were monitored using a UV–vis spectrophotometer
for the experiments. at the λmax of model pollutants (Table 1).

2.2. Photocatalyst preparation 2.5. Effect of light source temperature

In a typical experiment, 1.51 g of Bi(NO3)3⋅5H2O was firstly dis­ To understand the effect of light bulb color temperature on the
solved into 10 mL DMF and stirred for 45 min. 0.6435 g of KI was dis­ photocatalytic performance of 1-D BiOI, dye removal was studied using
solved into 5 mL dDW to prepare the second solution. The KI solution light sources with different temperatures. Incandescent, soft white light,
was then added to the Bi(NO3)3⋅5H2O solution dropwise under magnetic cool white light, and compact fluorescent bulbs were used as light
stirring, and the mixture was stirred further for 90 min at room tem­ irradiation sources. Dye degradation tests conducted on solutions of 25
perature. The obtained mixture was transferred to a 100 mL Teflon-lined mg/L EBT contained 400 mg/L photocatalysts. The dye degradation by
autoclave. The autoclave was heated in an oven at 130 ◦ C for 16 h. The photolysis was also checked to assure the photocatalysis degradation is
obtained solids were collected by centrifugation, washed several times responsible for dye removal. The degradation efficiency, DE%, of model
with dDW and ethanol, and dried at 60 ◦ C for 6 h (Fig. 1). pollutants was calculated using the following equation:
(Co − Ct )
2.3. Photocatalyst characterization DE% = × 100 (1)
Co

The crystal structure of the obtained photocatalyst was evaluated where Co is the initial concentration of organic dyes and Ct is the con­
using an X-ray diffractometer (XRD, Unisantis S.A, XMD300 model, centration of organic dyes at time t.
Geneva, Switzerland) with monochromatic Cukα (λ = 0.15406 nm) as a
radiation source. Fourier transform infrared (FT-IR) spectrum was
evaluated in the range from 4000 to 400 cm− 1 by using FT-IR, Thermo 2.6. Kinetic study
Nicolet, Avatar 370 instrument. The UV–vis diffuse reflection spectros­
copy (DRS) of the prepared 1-D BiOI was recorded on a UV–vis spec­ Kinetic models i.e. first-order and second-order models were used to
trophotometer (Unico UV-2100). The photoluminescence (PL) spectrum determine the rate of dye degradation. Firstly, 400 mg/L 1-D BiOI
was accomplished using Varian Cary Eclipse Fluorescence Spectropho­ photocatalysts were dispersed into EBT solutions with different con­
tometer. Scanning electron microscopy (SEM, MIRA3 TESCAN, Czech centrations of 25, 50, and 75 mg/L, respectively. Before irradiation, the
Republic) was used to analyze the surface morphology of the obtained solutions were agitated for 30 min in dark to establish adsorption-
photocatalyst. The specific surface area, porosity, and pore size distri­ desorption equilibrium on the catalyst surface. The samples then irra­
bution of the as-synthesized sample were measured by N2 adsorption- diated by a visible light source for 10, 20, 30, 45, 60, and 90 min.
desorption isotherms at 77 K according to the Brunauer–Emmett–­ Finally, the photocatalyst was separated through centrifugation and the
Teller (BET) and Barrett-Joyner-Halenda (BJH) methods (ELSORP-mini- dye concentration in supernatant was measured with a UV–vis spec­
II (BEL Japan, Inc.). trophotometer at λmax of EBT (530 nm).

Table 1
The chemical structure and properties of the present azo dyes.
Molecule Structure Formula Dye type λmax (nm)

C14H14N3NaO3S
Methyl Orange (MO) Anionic 465

Methylene Blue (MB) C16H18ClN3S Cationic 665

Eriochrome Black T (EBT) C20H12N3O7SNa Anionic 530

3
M. Nourzad et al. Catalysis Communications 174 (2023) 106567

Fig. 1. The schematic of 1-D BiOI preparation method.

2.7. Experimental design and optimization 2.8. Reusability of photocatalyst

Box-Behnken Design (BBD), the most widely used form of response The reusability of photocatalysts is an imperative parameter in the
surface method (RSM) was applied for modeling, optimizing, and economic viability of the process. Herein, the performance of the pho­
determining the effect of independent variables and the interactions on tocatalyst in consecutive use/ reuse experiments was performed in three
the response (EBT dye removal). Design-Expert (version 12) software cycles under optimal conditions. After completing the photocatalysis, 1-
was used to design a BBD for 4 independent variables i.e. pH, the dose of D BiOI was separated and dried at 70 ◦ C for the subsequent photo­
photocatalyst (g/L), irradiation time (min), and dye concentration (mg/ catalysis cycles.
L). The factors have been examined at three levels as listed in Table 2.
The variables were coded according to the following equation: 2.9. Chemical oxygen demand analysis
x − xo
α= i (2)
Δx The COD is a standard parameter used to measure the level of organic
matters in water. COD tests were used to measure dye calcination into
where α is the coded value of each independent variable; xi is the cor­ CO2, H2O, and other inorganics. The degree of EBT mineralization in
responding real value; x0 is the value at the center point, and Δx is the optimum operation conditions was calculated by the difference between
step-change in the variable xi. Based on Eq. (3), the design matrix was the initial COD of dye solution with 10 and 25 mg/L EBT (COD0), and
composed of a total of 27 runs (N) investigated experimentally. In the the remaining COD after photocatalysis (CODt) in Eq. 5.
equation, k is the number of factors and Cp is the number of replicates in
the central points: COD removal (%) =
COD0 − CODt
× 100 (5)
COD0
N = 2k (k − 1) + Cp (3)

A second-order model (Eq. 4) was used to establish the correlation 2.10. Mechanism study
between four independent variables and EBT degradation performance.
∑4 ∑4 ∑4 To understand the major active species generated by 1-D BiOI and
Y (%) = β0 + β Xi +
i=1 i
β X2 +
i=1 ii i
β Xi Xj
i<j ij
(4) contribute to the degradation of azo dye, radical scavengers were
applied to solutions. Environmental conditions were adjusted to the
where Y represents the EBT degradation efficiency; Xi is the independent optimum values obtained in the optimization step. A concentration of
variables; and (β0, βi, βii, and βij) are the model terms that stand for the 0.2 mol/L of AgNO3, KI, AA, and IPA was applied as scavengers to
intercept, quadratic, linear, and interaction effects, respectively. The quench e− , .OH, .O−2, and h+, respectively.
essential regression parameters such as the correlation coefficient (R2),
adjusted R2, p-value, and lack-of-fit (LOF) value were estimated by using 3. Results and discussion
analysis of variance (ANOVA) and used for the determination of the
relevancy and suitability of the predicated model. 3.1. Characterization of 1-D BiOI

The XRD pattern of 1-D BiOI is depicted in Fig. 2a. The main peaks of
1-D BiOI are observed at 29.7, 31.8, 45.5, and 55.2o, which can be
Table 2
Range and levels of independent variables used in BBD designs matrix. indexed to the (102), (110), (200), and (212) crystal planes, respec­
tively, characteristic for BiOI structures. The sharp diffraction peaks
Factor Variable level
indicated the high crystallinity of 1-D BiOI. The average crystallite size,
Code − 1 0 +1 Dhkl, of the 1-D BiOI was estimated using the Scherrer formula:
pH A 4 7 10
k×λ
1-D BiOI dose (g /L) B 0.1 0.55 1 Dhkl = (6)
Irradiation time (min) C 15 52.5 90 Bhkl cosθhkl
Dye (mg/L) D 10 30 50
where λ is the X-ray wavelength, β is the full width at half maximum, θ is
the Bragg angle, and k is the so-called shape factor which is equal to 0.9.
The average crystallite size of 1-D BiOI was calculated as 20.4 nm.

4
M. Nourzad et al. Catalysis Communications 174 (2023) 106567

Fig. 2. The powder XRD pattern (a) and the FT-IR spectrum of the as-synthesized 1-D BiOI sample (b).

Fig. 2b shows the FT-IR spectrum of the prepared 1-D BiOI. The EVB = ECB + Eg (9)
bands at 494.4 and 758.4 cm− 1 are assigned to the symmetric stretching
vibration of the Bi–O bond and the bands observed at 547.1, 1631.5, where ECB and EVB are potential energy (eV) of the CB and VB, respec­
and 3367.7 cm− 1 are attributed to the vibrations of the I–O bond, tively. X is the absolute electronegativity of a semiconductor, and Ee is
δ(O–H) bending, and stretching vibrations of ν(O–H) respectively. The the energy of free electrons in a standard hydrogen electrode (4.5 eV).
most intense band with a maximum at 1531.1 cm− 1 is attributed to the The Eg and X of 1-D BiOI were calculated as 1.90 and 5.99 eV, respec­
stretching vibration of the tetragonal crystal links BiOI [62]. tively. ECB and EVB were calculated at 0.54 and 2.44 eV, respectively.
The optical property of 1-D BiOI was evaluated based on the diffuse The PL spectrum of the 1-D BiOI under an excitation wavelength of
reflectance spectrum (DRS) in UV–vis region (Fig. 3a) and the photo­ 265 nm revealed three emission peaks located at 264, 420, and 533 nm,
luminescence (PL) spectrum (Fig. 3b). respectively.
As shown in Fig. 3a, 1-D BiOI has an intense visible light absorption The morphology of the prepared 1-D BiOI was scanned by SEM
efficiency with an absorption edge at about 653 nm, a fact in good analysis (Fig. 4). It can be observed that 1-D BiOI consists of BiOI
correlation with the bright red color of the 1-D BiOI powder. The nanosheets with an average diameter of <500 nm, assembled to form a
bandgap of the 1-D BiOI was calculated either based on the extrapola­ flower-like micro-spherical structure.
tion in the Tauc plot or based on the simplified equation: The N2 adsorption-desorption isotherms of the 1-D BiOI are illus­
1240 trated in Fig. 5a. The specific surface area based on BET method (SBET)
Eg = (7) of 1-D BiOI was found 32 m2/g, while the total pore volume (VTot)
λ
0.193 cm3/g. As shown in Fig. 5a, the N2 isotherm is closer to Type IV
where λ is the wavelength and Eg is the bandgap. The Eg of 1-D BiOI was type with an H3 type hysteresis [65]. This suggested that the prepared
calculated as 1.90 eV, a value in line with the reported ones in the 1-D BiOI can be assumed as non-porous, with the measured volume
literature [62]. of mesoporous with a wide range of size distribution (Fig. 5b) to
The valence band (VB) and conduction band (CB) potentials of the 1- be linked to the formed inter-particles, inter-clusters and inter-
D BiOI were calculated by the following equations [63,64]: aggregates and spaces [66,67] as a result of the inter-stacking of the
nanosheets.
ECB = X − Ee − 0.5Eg (8)

Fig. 3. The UV–vis DRS spectrum (a) and the PL spectrum of the 1-D BiOI at 265 nm excitation wavelength (b).

5
M. Nourzad et al. Catalysis Communications 174 (2023) 106567

Fig. 4. SEM images of pristine 1-D BiOI at different magnifications.

Fig. 5. The N2 adsorption-desorption isotherm (a) and the pore size distribution based on the BJH method (b) of 1-D BiOI.

3.2. Photocatalytic performance for model contaminants investigated. As shown in Fig. 6, the maximum photo-assisted removal
percentages for MO, MB, and EBT were 47.4 and 29.2%, 66.1 and
To determine the degradation performance of 1-D BiOI for model azo 51.1%, and 86.7 and 84.6% for solutions of 12 and 15 mg/L dye,
dyes, the removal of MO, MB, and EBT at two different initial concen­ respectively. Interestingly, in the case of EBT, the photo-assisted
trations, and under irradiation by a 15 W soft white light bulb was removal was significantly higher compared to the other two dyes and
less dependent to the increment of the initial dye concentration,
revealing that the material can perform well even for higher concen­
trations. More research effort should be devoted to this aspect since it
was outside of the scope of this project. Based on the above preliminary
results, we focused on studying in more detail the EBT remediation and
optimizing various parameters.

3.3. Effect of light source temperature

Since the highest preliminary photo-assisted removal was found


against the azo dye EBT, we focused as mentioned and above on its
removal with emphasis also on how different sources of light can affect
the high photo-assisted removal efficiency.
As a first step, we explored the photo-stability of the dyes based on
photolysis tests (absence of photocatalyst) with the results to be
collected in Fig. 7a. As it can be observed, EBT showed elevated photo-
stability in all cases, since even after 120 min of light exposure <5%
photolysis/conversion was recorded, except when the fluorescent lamp
Fig. 6. Photo-assisted removal of three azo dyes (Methyl Orange (MO),
Methylene Blue (MB), and Eriochrome Black T (EBT)) at two different initial was used, with the photolysis reaching around 9% after 120 min.
concentrations of 12 mg/L (right bar) and 15 mg/L (left bar), (pH: 7.5, 1-D Regarding the photocatalytic-assisted conversion of EBT, 30 min in
BiOI: 400 mg/L, light source:15 W soft white light bulb, irradiation time: 60 the dark equilibration was found sufficient in the case of 25 mg/L initial
min). (For interpretation of the references to color in this figure legend, the EBT concentration. The adsorption efficiency of 1-D BiOI was recorded
reader is referred to the web version of this article.) elevated (Fig. 7b) since almost half of the azo-dye molecules were

6
M. Nourzad et al. Catalysis Communications 174 (2023) 106567

Fig. 7. Photolysis of EBT dye (a) and the degradation of EBT dye as a function of the visible light source (pH: 7.5, 1-D BiOI: 400 mg/L, EBT: 25 mg/L) (b).

adsorbed during the in-the-dark stabilization and hence the adsorption degradation as absolute percentage values decreased from 96 to 64% by
capacity was around 63 mg/g. Upon light irradiation, it can be observed increasing dye concentration from 25 to 75 mg/L. Although, it should be
that all four lamps were capable to “turn-on” the material’s photo­ considered that the amount of dye removed in the case of 75 mg/L initial
catalytic activity. The smoother kinetics, as well as the highest removal concentration is double compared to the one removed in the case of 25
performance, were shown in the case of 15 W soft white light bulb, mg/L. Fig. 8b and c illustrated the application of the kinetic models on
hence this source of light was chosen for further investigations/experi­ EBT dye degradation using different initial dye concentrations. From the
ments in order to optimize the process and to determine which param­ linear fitting curves of ln (C/Co) and (1/C)-(1/CO) versus irradiation
eters are important. time calculated the EBT degradation rate constants, kapp, and R2. Table 3
gives the kapp and R2 values for EBT photodegradation by 1-D BiOI
3.4. Photocatalytic degradation kinetic which indicated that the first-order model fits better compared to the
second-order model for the different dye concentrations. Furthermore,
The kinetic of EBT degradation was studied by analyzing the dye
removal using different initial dye concentrations, specifically of 25, 50, Table 3
and 75 mg/L. Two common kinetic models, the first-order and second- The kinetic constants and R2 for different kinetic models.
order, were used to model the photo-assisted degradation kinetics.
Dye Max First-order Second-order
The first-order and second-order kinetic models are given in Eq. (10) and concentration Removal
kapp R2 kapp R2
Eq. (11), respectively: (mg/L) (%)
(min− 1) (L mg− 1
C min− 1)
ln = − kapp t (10)
C0 25 96 0.02872 0.97 0.00353 0.35
± 0.0013 ± 0.00195
1 1 50 68 0.0093 0.92 2.44447 × 0.84
− = kapp t (11) ± 0.00113 10− 4
C C0
± 4.29892 ×
10− 5
where kapp is the rate constant. C0 and C are the concentration of dye at 75 64 0.00886 0.96 1.65418 × − 4 0.78
initial (t = 0) and certain reaction time (t = t). ± 7.27649 × ± 3.5805 ×
Fig. 8a shows the effect of initial dye concentration on EBT photo- 10− 4 10− 5
assisted removal evolution. The results reveal that EBT dye

Fig. 8. The evolution of photo-assisted remediation by time (a), first-order kinetics (b), and second-order kinetics (c) using three different initial EBT concentrations
(pH: 7.5, 1-D BiOI: 400 mg/L, light source: 15 W soft white light bulb).

7
M. Nourzad et al. Catalysis Communications 174 (2023) 106567

and as expected, the kapp constant depends on the dye concentration and
decreases with increasing Co. A very crucial outcome is that the photo­
catalytic EBT decomposition is negatively affected by increasing the
initial concentration, a fact that suggests the optimization of various
parameters plays a key role.

3.5. Mathematical modeling and optimization

3.5.1. Model fitting and statistical analysis


Box-Behnken Design (BBD) and modeling were used to determine the
effects of four independent variables and their effect on EBT degradation
in order to conclude the optimum conditions/parameters. The BBD
design matrix and responses (EBT removal) are presented in Table 4.
ANOVA was used to determine the best fit model for observed
(experimental) data in Table 4. The results of fitting models are shown in
Table S1. The F value and P value were used as tools to determine the
best fit model and the significant variables in EBT removal. Table S1
Fig. 9. Experimental vs. predicted photocatalytic EBT degradation by 1-D BiOI.
clearly shows that the quadratic model gives the best description for dye
degradation data. ANOVA for the quadratic model fitted on experi­
mental data is presented in Table S2. A P value<0.05 and a large F value 3.5.2. Effect of operating variables and optimization
indicated the significant role of the variable in the process. Moreover, 3D surface plots given in Fig. 10(a–c) were used to visualize the ef­
the coefficient of determination (R2) and LOF test were used to reflect fects of various studied variables on EBT degradation efficiencies by 1-D
the model adequacy. BiOI. Fig. 10a shows the mutual effects of photocatalyst dose and initial
For a statistically adequate model, R2 values are nearly 1, LOF > EBT concentration on EBT degradation. As seen, dye degradation
0.05, R2pred, and R2adj are within ±0.2. As seen, R2, R2pred, and R2adj are increased directly by catalyst dose in the range of 0.1 to about 1 g/L.
0.94, 0.69, and 0.88, respectively which indicates the quadratic model Moreover, the rate of dye degradation increased sharply by 1-D BiOI
describes the data very accurately. Moreover, the adequate precision dose in the range of 0.1–0.55 g/L, and then continued slightly in the
which measures the signal-to-noise ratio is 15.43 which is above the range of 0.55–1 g/L. Decreasing the rate of EBT removal at a higher 1-D
minimum desirable of 4. Based on Table S2, A, B, C, D, B2, and C2 are BiOI dose could be attributed to the inhibition of light penetration in a
statistically significant terms in the process. ANOVA for the coded fac­ turbid solution caused by the concentrated catalyst. Limiting light
tors gives a polynomial mathematical equation that describes the EBT penetration also underlies the lower dye removal at the higher EBT
degradation process. The equation could be used to predict the dye concentrations. Furthermore, a higher dye removal at dilute solutions
removal percentage under specified environmental conditions. The could be attributed to the higher chance of dye degradation by limited
model in terms of coded factors is presented in Eq. (12): active species in the solution.
Fig. 10b shows the effect of light exposure time and photocatalyst
EBT removal (%) = 64.6 + 17.4A + 17.4B + 8.05C − 21.6D + 1.1AB dose on the dye degradation efficiency. As can be seen, dye removal
− 2.1175 AC − 2.2575 CE − 3.3 BCE + .2.9BD increased proportionally by irradiation time in the range of 15–90 min.
− 7.1CD + 5.8A2 –11.0B2 –9.4C2 –1.9D2 This is a common behavior in photocatalysis that is due to the higher
time EBT molecules have to be degraded by active radicals. As is evident
(12)
in Fig. 10c, dye degradation efficiency increases with increasing pH from
where A, B, C, and D represent coded values of pH, the dose of photo­ 4 to 10, with the highest removal to be found at pH 10.
catalyst, irradiation time, and initial EBT concentration, respectively. Optimization is the final goal in RSM modeling and aimed to figure
Fig. 9 visualized the experimental vs. predicted data for the photo­ out the parametric condition in which the highest EBT removal occurs.
catalytic degradation of EBT by 1-D BiOI. As seen, the experimental data Optimum levels for the highest degradation were explored by solving
were distributed close to their corresponding predicted values by the Eq. 4 in the studied range of independent variables in Table 2. As shown
model which indicates a high precision of the model. in Table 5, the highest 99.4% EBT degradation was predicted to occur
when pH, reaction time, dose, and EBT concentration were adjusted at

Table 4
BBD design matrix for EBT degradation by 1-D BiOI.
Run No Coded Response Run No Coded Response
variable (% removal) variable (% removal)

A B C D Observed Predicted A B C D Observed predicted

1 0 − 1 1 0 31.7 38.2 15 0 − 1 0 1 10.2 9.7


2 − 1 0 − 1 0 36 33.4 16 0 0 − 1 − 1 68.4 59.8
3 1 0 1 0 87.1 84.4 17 0 0 1 − 1 98.9 90.1
4 0 0 1 1 33 32.6 18 0 0 0 0 62.7 64.6
5 0 − 1 0 − 1 62.7 58.8 19 1 0 0 − 1 99 >100
6 0 1 0 1 51.8 50.3 20 0 0 0 0 68.0 64.6
7 − 1 0 1 0 58 53.7 21 0 1 0 − 1 92.6 87.8
8 0 1 − 1 0 49.3 57.0 22 − 1 − 1 0 0 30.4 25.7
9 1 1 0 0 99.5 95.3 23 − 1 1 0 0 65.2 58.3
10 1 0 0 1 63 62.0 24 0 -1 -1 0 10.7 15.4
11 -1 0 0 1 28.2 31.6 25 0 0 0 0 63.1 64.6
12 1 -1 0 0 60.4 58.4 26 0 1 1 0 56.8 66.3
13 1 0 -1 0 73.6 72.6 27 -1 0 0 -1 55.2 70.4
14 0 0 -1 1 30.9 30.7

8
M. Nourzad et al. Catalysis Communications 174 (2023) 106567

Fig. 10. 3D illustration of photocatalytic EBT degradation by 1-D BiOI as a function of different operating parameters. The photocatalyst dose and initial EBT
concentration (a) reaction time and photocatalyst dose (b) photocatalyst dose and solution pH (c).

Table 5
Optimum values were obtained for study variables in EBT degradation.
Parameter pH Reaction Dose EBT Removal (%)
time dye

Unit – min g/L mg/L Experimental Predicted

Optimum 7.97 76.64 0.724 10.27 98.1 99.4


value

7.97, 76.64 min, 0.724 g/L, and 10.27 mg/L, respectively. Validation
tests were experimented to ensure that the optimum condition suggested
is adequately precise. As shown in the table, the average dye degrada­
tion in the validity tests by three replicates was 98.1% which is
reasonably close to the predicted 99.4%.

3.6. Photocatalyst reusability

The reusability of the catalysts is a significant factor in the design and Fig. 11. The reusability of 1-D BiOI in consecutive cycles of photocatalytic
development of an environmentally friendly, sustainable, and cost- degradation (the experiments were performed at least twice with the error
effective process. Hence, the reusability of 1-D BiOI was evaluated in a deviation to be <6%).
series of successive use-reuse cycles for EBT degradation, and the results
are presented in Fig. 11. As seen, EBT removal efficiency decreased quite insignificant loss in the dye photocatalytic degradation efficiency during
slightly from 97.0% for the pristine material to 92.1% after three cycles. the experiments and the structural stability of 1-D BiOI reflected the
Furthermore, the structural properties of exhausted 1-D BiOI after advantages of 1-D BiOI for practical application.
three cycles were analyzed by SEM and the results are presented in
Fig. 12. As can be observed, no obvious morphological destruction was
observed in nanosheets of micro-spherical 1-D BiOI structure. The

9
M. Nourzad et al. Catalysis Communications 174 (2023) 106567

Fig. 12. SEM images of the reused 1-D BiOI at different magnifications.

Fig. 14. Effects of different scavengers on the EBT degradation over the 1-D
BiOI photocatalyst.

4. Conclusion

Flower-like spherical aggregations consisting of nano 1-D BiOI with a


specific surface area of 32 m2/g were successfully synthesized via a
simple solvothermal method and utilized as efficient photo-assisted
remediation material against different azo dyes. Investigation of irra­
diation by different light sources showed that the low-intensity soft
white light bulb activated the catalyst excellently for EBT degradation.
The first-order model described the EBT degradation well. A mathe­
Fig. 13. COD removal during the EBT photocatalytic degradation (pH: 7.97, 1-
matical model was developed by performing the study according to BBD.
D BiOI: 0.724 g/L, light source: 15 W soft white light bulb).
The model optimization predicted the highest photo-degradation of
98.1% using 0.724 g/L 1-D BiOI and 76.64 min irradiation. Radical
3.7. Chemical oxygen demand (COD) tests
trapping/scavenging experiments indicated that •OH constitutes the
main reactive species for EBT degradation. As an indication of dye
COD tests were used to investigate the degree of EBT dye calcination
calcination, the COD of dye solutions increased first during the irradi­
during photocatalysis. Herein, the COD of dye solutions at two different
ation time and then decreased as the dye further mineralized. Reuse tests
concentrations (10 and 25 mg/L) of EBT was monitored for different
for 1-D BiOI proved the premium reusability of 1-D BiOI with a negli­
irradiation times. The results given in Fig. 13 show a sharp increase in
gible loss in removal efficacy.
COD values after about 20 min of irradiation. The levels of COD in the
solutions then decreased gradually and reached a minimum value,
CRediT authorship contribution statement
where the majority of organic molecules were mineralized to inorganic
species.
Marzieh Nourzad: Investigation. Aliakbar Dehghan: Investigation,
Writing – original draft, Formal analysis. Zohreh Niazi: Writing –
3.8. Photodegradation mechanism
original draft, Formal analysis. Dimitrios A. Giannakoudakis: Writing
– original draft, Conceptualization. Mojtaba Afsharnia: Writing –
To identify the main active species that are involved in the photo­
original draft. Mariusz Barczak: Writing – review & editing. Ioannis
catalytic process, AgNO3, KI, AA, and IPA were used as scavengers to
Anastopoulos: Writing – review & editing. Konstantinos S. Tri­
quench e− , .OH, .O−2, and h+, respectively [68]. Fig. 14 displays the EBT
antafyllidis: Writing – review & editing. Mahmoud Shams: Writing –
degradation by 1-D BiOI in the presence of different radical capture
review & editing, Resources, Project administration, Conceptualization.
agents. EBT degradation by 1-D BiOI in the absence of scavengers is also
shown in the graph. As shown, EBT degradation was significantly sup­
Declaration of Competing Interest
pressed from 96.5% to 83.4% by adding KI. This inhibition implies that
the •OH radicals play a key role in the degradation of EBT. An insig­
The authors declare that they have no known competing financial
nificant decrease in EBT removal (91.4%) compared to the control
interests or personal relationships that could have appeared to influence
(96.5%) when IPA and AA were added to capture h+ and O•– 2 indicated
the work reported in this paper.
that these photoinduced species plays an insignificant role in the pho­
tocatalytic decomposition. Furthermore, when AgNO3 was introduced
Data availability
to capture electrons, the EBT degradation efficiency was increased to
96.7%, mainly because the capture of electrons reduces the chance of
No data was used for the research described in the article.
e− /h+ recombination. By quenching the e− /h+ recombination, the
number of electrons available to produce. O•– 2 radicals increases, and
Acknowledgements
more hydroxyl radicals are formed through splitting H2O molecules by
holes.
To Mashhad university of medical sciences for financial support of
the MSC dissertation under grant # of 991310.

10
M. Nourzad et al. Catalysis Communications 174 (2023) 106567

Appendix A. Supplementary data monodisperse photocatalyst, J. Solid State Chem. 300 (2021), 122287, https://doi.
org/10.1016/j.jssc.2021.122287.
[20] Y. Kaur, T. Jasrotia, R. Kumar, G.R. Chaudhary, S. Chaudhary, Adsorptive removal
Supplementary data to this article can be found online at https://doi. of eriochrome black T (EBT) dye by using surface active low cost zinc oxide
org/10.1016/j.catcom.2022.106567. nanoparticles: a comparative overview, Chemosphere. 278 (2021), 130366,
https://doi.org/10.1016/j.chemosphere.2021.130366.
[21] M. Coha, G. Farinelli, A. Tiraferri, M. Minella, D. Vione, Advanced oxidation
References processes in the removal of organic substances from produced water: potential,
configurations, and research needs, Chem. Eng. J. 414 (2021), 128668, https://doi.
[1] J.-H. Shen, Z.-W. Jiang, D.-Q. Liao, J.-J. Horng, Enhanced synergistic org/10.1016/j.cej.2021.128668.
photocatalytic activity of TiO2/oxidant for azo dye degradation under simulated [22] F.E. Titchou, H. Zazou, H. Afanga, J. El Gaayda, R. Ait Akbour, P.V. Nidheesh,
solar irradiation: a determination of product formation regularity by quantifying M. Hamdani, Removal of organic pollutants from wastewater by advanced
hydroxyl radical-reacted efficiency, J. Water Process Eng. 40 (2021), 101893, oxidation processes and its combination with membrane processes, Chem. Eng.
https://doi.org/10.1016/j.jwpe.2020.101893. Process. Process Intensif. 169 (2021), 108631, https://doi.org/10.1016/j.
[2] X. Zhang, J. Wang, X.-X. Dong, Y.-K. Lv, Functionalized metal-organic frameworks cep.2021.108631.
for photocatalytic degradation of organic pollutants in environment, Chemosphere. [23] M. Arumugam, T.S. Natarajan, T. Saelee, S. Praserthdam, M. Ashokkumar,
242 (2020), 125144, https://doi.org/10.1016/j.chemosphere.2019.125144. P. Praserthdam, Recent developments on bismuth oxyhalides (BiOX; X = cl, Br, I)
[3] D. Chen, Y. Cheng, N. Zhou, P. Chen, Y. Wang, K. Li, S. Huo, P. Cheng, P. Peng, based ternary nanocomposite photocatalysts for environmental applications,
R. Zhang, L. Wang, H. Liu, Y. Liu, R. Ruan, Photocatalytic degradation of organic Chemosphere. 282 (2021), 131054, https://doi.org/10.1016/j.
pollutants using TiO2-based photocatalysts: a review, J. Clean. Prod. 268 (2020), chemosphere.2021.131054.
121725, https://doi.org/10.1016/j.jclepro.2020.121725. [24] T. Bavani, J. Madhavan, S. Prasad, M.S. AlSalhi, M. ALJaffreh, S. Vijayanand,
[4] Y. Pi, X. Li, Q. Xia, J. Wu, Y. Li, J. Xiao, Z. Li, Adsorptive and photocatalytic Fabrication of novel AgVO3/BiOI nanocomposite photocatalyst with
removal of persistent organic pollutants (POPs) in water by metal-organic photoelectrochemical activity towards the degradation of rhodamine B under
frameworks (MOFs), Chem. Eng. J. 337 (2018) 351–371, https://doi.org/10.1016/ visible light irradiation, Environ. Res. 200 (2021), 111365, https://doi.org/
j.cej.2017.12.092. 10.1016/j.envres.2021.111365.
[5] S. Samsami, M. Mohamadi, M.H. Sarrafzadeh, E.R. Rene, M. Firoozbahr, Recent [25] Y.H. Lee, Y.M. Dai, J.Y. Fu, C.C. Chen, A series of bismuth-oxychloride/bismuth-
advances in the treatment of dye-containing wastewater from textile industries: oxyiodide/graphene-oxide nanocomposites: synthesis, characterization, and
overview and perspectives, Process. Saf. Environ. Prot. 143 (2020) 138–163, photcatalytic activity and mechanism, Mol. Catal. 432 (2017) 196–209, https://
https://doi.org/10.1016/j.psep.2020.05.034. doi.org/10.1016/j.mcat.2017.01.002.
[6] M. Baziar, H.R. Zakeri, Z.D. Nejad, M. Shams, I. Anastopoulos, D. [26] S.-Y. Chou, W.-H. Chung, L.-W. Chen, Y.-M. Dai, W.-Y. Lin, J.-H. Lin, C.-C. Chen,
A. Giannakoudakis, E.C. Lima, Metal-organic and Zeolitic imidazole frameworks as A series of BiO x I y /GO photocatalysts: synthesis, characterization, activity, and
cationic dye adsorbents: Physicochemical optimizations by parametric modeling mechanism, RSC Adv. 6 (2016) 82743–82758, https://doi.org/10.1039/
and kinetic studies, J. Mol. Liq. 332 (June 2021) (2021) 115832, https://doi.org/ C6RA12482H.
10.1016/j.molliq.2021.115832, 115832. [27] P. Dumrongrojthanath, T. Thongtem, A. Phuruangrat, S. Thongtem, Synthesis and
[7] S. Marimuthu, A.J. Antonisamy, S. Malayandi, K. Rajendran, P.-C. Tsai, characterization of hierarchical multilayered flower-like assemblies of ag doped
A. Pugazhendhi, V.K. Ponnusamy, Silver nanoparticles in dye effluent treatment: a Bi2WO6 and their photocatalytic activities, Superlattice. Microst. 64 (2013)
review on synthesis, treatment methods, mechanisms, photocatalytic degradation, 196–203, https://doi.org/10.1016/j.spmi.2013.09.028.
toxic effects and mitigation of toxicity, J. Photochem. Photobiol. B Biol. 205 [28] O. Yayapao, T. Thongtem, A. Phuruangrat, S. Thongtem, Ultrasonic-assisted
(2020), 111823, https://doi.org/10.1016/j.jphotobiol.2020.111823. synthesis of Nd-doped ZnO for photocatalysis, Mater. Lett. 90 (2013) 83–86,
[8] S. Benkhaya, S.M’. Rabet, A. El Harfi, A review on classifications, recent synthesis https://doi.org/10.1016/j.matlet.2012.09.027.
and applications of textile dyes, Inorg. Chem. Commun. 115 (2020), 107891, [29] O. Yayapao, T. Thongtem, A. Phuruangrat, S. Thongtem, Sonochemical synthesis of
https://doi.org/10.1016/j.inoche.2020.107891. Dy-doped ZnO nanostructures and their photocatalytic properties, J. Alloys
[9] R. Kishor, D. Purchase, G.D. Saratale, L.F.R. Ferreira, M. Bilal, H.M.N. Iqbal, R. Compd. 576 (2013) 72–79, https://doi.org/10.1016/j.jallcom.2013.04.133.
N. Bharagava, Environment friendly degradation and detoxification of Congo red [30] J. Akhtar, M.B. Tahir, M. Sagir, H.S. Bamufleh, Improved photocatalytic
dye and textile industry wastewater by a newly isolated Bacillus cohnni (RKS9), performance of Gd and Nd co-doped ZnO nanorods for the degradation of
Environ. Technol. Innov. 22 (2021), 101425, https://doi.org/10.1016/j. methylene blue, Ceram. Int. 46 (2020) 11955–11961, https://doi.org/10.1016/j.
eti.2021.101425. ceramint.2020.01.234.
[10] D. Sivaraj, K. Vijayalakshmi, M. Srinivasan, P. Ramasamy, Graphene oxide [31] H. Mittal, M. Khanuja, Hydrothermal in-situ synthesis of MoSe2-polypyrrole
reinforced bismuth titanate for photocatalytic degradation of azo dye (DB15) nanocomposite for efficient photocatalytic degradation of dyes under dark and
prepared by hydrothermal method, Ceram. Int. 47 (2021) 25074–25080, https:// visible light irradiation, Sep. Purif. Technol. 254 (2021), 117508, https://doi.org/
doi.org/10.1016/j.ceramint.2021.05.238. 10.1016/j.seppur.2020.117508.
[11] G. Palanisamy, T. Pazhanivel, K. Bhuvaneswari, G. Bharathi, G. Marimuthu, [32] Y.C. Chou, Y.Y. Lin, C.S. Lu, F.Y. Liu, J.H. Lin, F.H. Chen, C.C. Chen, W.T. Wu,
T. Maiyalagan, Spinel oxide ZnCr2O4 incorporated with ZnS quantum dots for Controlled hydrothermal synthesis of BiOxCly/BiOmBrn/g-C3N4 composites
application on visible light driven photocatalyst azo dye degradation, Colloids exhibiting visible-light photocatalytic activity, J. Environ. Manag. 297 (2021),
Surfaces A Physicochem. Eng. Asp. 590 (2020), 124505, https://doi.org/10.1016/ 113256, https://doi.org/10.1016/j.jenvman.2021.113256.
j.colsurfa.2020.124505. [33] C.C. Chen, J.Y. Fu, J.L. Chang, S.T. Huang, T.W. Yeh, J.T. Hung, P.H. Huang, F.
[12] M. Yousefi, M. Gholami, V. Oskoei, A.A. Mohammadi, M. Baziar, A. Esrafili, Y. Liu, L.W. Chen, Bismuth oxyfluoride/bismuth oxyiodide nanocomposites
Comparison of LSSVM and RSM in simulating the removal of ciprofloxacin from enhance visible-light-driven photocatalytic activity, J. Colloid Interface Sci. 532
aqueous solutions using magnetization of functionalized multi-walled carbon (2018) 375–386, https://doi.org/10.1016/j.jcis.2018.07.130.
nanotubes: Process optimization using GA and RSM techniques, J. Environ. Chem. [34] J.Y. Fu, L.W. Chen, Y.M. Dai, F.Y. Liu, S.T. Huang, C.C. Chen, BiOmFn/BiOxIy/GO
Eng 9 (4) (2021) 105677, https://doi.org/10.1016/j.jece.2021.105677, 105677. nanocomposites: synthesis, characterization, and photocatalytic activity, Mol.
[13] M.M. Haque, M.A. Haque, M.K. Mosharaf, P.K. Marcus, Decolorization, Catal. 455 (2018) 214–223, https://doi.org/10.1016/j.mcat.2018.06.014.
degradation and detoxification of carcinogenic sulfonated azo dye methyl orange [35] C.W. Siao, H.L. Chen, L.W. Chen, J.L. Chang, T.W. Yeh, C.C. Chen, Controlled
by newly developed biofilm consortia, Saudi, J. Biol. Sci. 28 (2021) 793–804, hydrothermal synthesis of bismuth oxychloride/bismuth oxybromide/bismuth
https://doi.org/10.1016/j.sjbs.2020.11.012. oxyiodide composites exhibiting visible-light photocatalytic degradation of 2-
[14] H. Chandarana, P. Senthil Kumar, M. Seenuvasan, M. Anil Kumar, Kinetics, hydroxybenzoic acid and crystal violet, J. Colloid Interface Sci. 526 (2018)
equilibrium and thermodynamic investigations of methylene blue dye removal 322–336, https://doi.org/10.1016/j.jcis.2018.04.097.
using Casuarina equisetifolia pines, Chemosphere. 285 (2021), 131480, https:// [36] J. Liu, Z. Luo, W. Han, Y. Zhao, P. Li, Materials science in semiconductor processing
doi.org/10.1016/j.chemosphere.2021.131480. preparation of ZnO / Bi 2 WO 6 heterostructures with improved photocatalytic
[15] M. Najjar, H.A. Hosseini, A. Masoudi, Z. Sabouri, A. Mostafapour, M. Khatami, performance, Mater. Sci. Semicond. Process. 106 (2020), 104761.
M. Darroudi, Green chemical approach for the synthesis of SnO2 nanoparticles and [37] Z. Liu, K. Xu, H. Yu, M. Zhang, Z. Sun, In-situ preparation of double Z-scheme
its application in photocatalytic degradation of Eriochrome black T dye, Optik Bi2S3/BiVO4/TiO2 ternary photocatalysts for enhanced photoelectrochemical and
(Stuttg). 242 (2021) 1–13, https://doi.org/10.1016/j.ijleo.2021.167152. photocatalytic performance, Appl. Surf. Sci. 545 (2021) 1–9, https://doi.org/
[16] A. Raza, M. Ikram, M. Aqeel, M. Imran, A. Ul-Hamid, K.N. Riaz, S. Ali, Enhanced 10.1016/j.apsusc.2021.148986.
industrial dye degradation using co doped in chemically exfoliated MoS2 [38] C.-C. Chen, S.-H. Chang, J. Shaya, F.-Y. Liu, Y.-Y. Lin, L.-G. Wang, H.-Y. Tsai, C.-
nanosheets, Appl. Nanosci. 10 (2020) 1535–1544, https://doi.org/10.1007/ S. Lu, Hydrothermal synthesis of BiOxBry/BiOmIn/GO composites with visible-
s13204-019-01239-3. light photocatalytic activity, J. Taiwan Inst. Chem. Eng. 133 (2022), 104272,
[17] M. Hasanpour, M. Hatami, Photocatalytic performance of aerogels for organic dyes https://doi.org/10.1016/j.jtice.2022.104272.
removal from wastewaters: review study, J. Mol. Liq. 309 (2020), 113094, https:// [39] Y.-Y. Lin, K.-Y. Hung, F.-Y. Liu, Y.-M. Dai, J.-H. Lin, C.-C. Chen, Photocatalysts of
doi.org/10.1016/j.molliq.2020.113094. quaternary composite, bismuth oxyfluoride/bismuth oxyiodide/ graphitic carbon
[18] O. Fawzi Suleiman Khasawneh, P. Palaniandy, Removal of organic pollutants from nitride: synthesis, characterization, and photocatalytic activity, Mol. Catal. 528
water by Fe2O3/TiO2 based photocatalytic degradation: a review, Environ. (2022), 112463, https://doi.org/10.1016/j.mcat.2022.112463.
Technol. Innov. 21 (2021), 101230, https://doi.org/10.1016/j.eti.2020.101230. [40] Y.-Y. Lin, J.-T. Hung, Y.-C. Chou, S.-J. Shen, W.-T. Wu, F.-Y. Liu, J.-H. Lin, C.-
[19] N.T. Tran, L.G. Trung, M.K. Nguyen, The degradation of organic dye contaminants C. Chen, Synthesis of bismuth oxybromochloroiodide/graphitic carbon nitride
in wastewater and solution from highly visible light responsive ZIF-67 quaternary composites (BiOxCly/BiOmBrn/BiOpIq/g-C3N4) enhances visible-

11
M. Nourzad et al. Catalysis Communications 174 (2023) 106567

light-driven photocatalytic activity, Catal. Commun. 163 (2022), 106418, https:// and application of visible-light-driven photocatalytic activities, J. Colloid Interface
doi.org/10.1016/j.catcom.2022.106418. Sci. 544 (2019) 25–36, https://doi.org/10.1016/j.jcis.2019.02.067.
[41] M. Yadav, S. Garg, A. Chandra, K. Hernadi, Immobilization of green BiOX (X= cl, [55] J. Hasan, G. Ouyang, J. Wang, H. Li, G. Tian, C. Qin, Efficient visible-light-driven
Br and I) photocatalysts on ceramic fibers for enhanced photocatalytic degradation photocatalysis of flower-like composites of AgI nanoparticle dotting BiOI
of recalcitrant organic pollutants and efficient regeneration process, Ceram. Int. 45 nanosheet, J. Solid State Chem. 297 (2021), 122044, https://doi.org/10.1016/j.
(2019) 17715–17722, https://doi.org/10.1016/j.ceramint.2019.05.340. jssc.2021.122044.
[42] Q. Yuan, S. Wei, T. Hu, Y. Ye, Y. Cai, J. Liu, P. Li, C. Liang, Defect-modified [56] X. Yang, Z. Chen, W. Zhao, C. Liu, X. Qian, W. Chang, T. Sun, C. Shen, G. Wei,
ultrathin BiOX (X = cl, Br) Nanosheets via a top-down approach with effective Construction of porous-hydrangea BiOBr/BiOI n-n heterojunction with enhanced
visible-light photocatalytic degradation, J. Phys. Chem. C 125 (2021) photodegradation of tetracycline hydrochloride under visible light, J. Alloys
18630–18639, https://doi.org/10.1021/acs.jpcc.1c02950. Compd. 864 (2021), 158784, https://doi.org/10.1016/j.jallcom.2021.158784.
[43] H. Liu, C. Yang, J. Huang, J. Chen, J. Zhong, J. Li, Ionic liquid-assisted [57] A.H. Lee, Y.C. Wang, C.C. Chen, Composite photocatalyst, tetragonal lead bismuth
hydrothermal preparation of BiOI/BiOCl heterojunctions with enhanced separation oxyiodide/bismuth oxyiodide/graphitic carbon nitride: synthesis, characterization,
efficiency of photo-generated charge pairs and photocatalytic performance, Inorg. and photocatalytic activity, J. Colloid Interface Sci. 533 (2019) 319–332, https://
Chem. Commun. 113 (2020), 107806, https://doi.org/10.1016/j. doi.org/10.1016/j.jcis.2018.08.008.
inoche.2020.107806. [58] K. Ren, K. Zhang, J. Liu, H. Luo, Y. Huang, X. Yu, Controllable synthesis of hollow/
[44] W. Jiang, Z. Li, C. Liu, D. Wang, G. Yan, B. Liu, G. Che, Enhanced visible-light- flower-like BiOI microspheres and highly efficient adsorption and photocatalytic
induced photocatalytic degradation of tetracycline using BiOI/MIL-125(Ti) activity, CrystEngComm 14 (2012) 4384–4390, https://doi.org/10.1039/
composite photocatalyst, J. Alloys Compd. 854 (2021), 157166, https://doi.org/ C2CE25087J. In press.
10.1016/j.jallcom.2020.157166. [59] Z. Jiang, X. Liang, Y. Liu, T. Jing, Z. Wang, X. Zhang, X. Qin, Y. Dai, B. Huang,
[45] Y. Guan, J. Wu, X. Man, Q. Liu, Y. Qi, P. He, X. Qi, Rational fabrication of flower- Enhancing visible light photocatalytic degradation performance and bactericidal
like BiOI1-x photocatalyst by modulating efficient iodine vacancies for mercury activity of BiOI via ultrathin-layer structure, Appl. Catal. B Environ. 211 (2017)
removal and DFT study, Chem. Eng. J. 396 (2020), 125234, https://doi.org/ 252–257, https://doi.org/10.1016/j.apcatb.2017.03.072.
10.1016/j.cej.2020.125234. [60] A. Dehghan, M.H. Dehghani, R. Nabizadeh, N. Ramezanian, M. Alimohammadi, A.
[46] M. Ji, J. Di, Y. Liu, R. Chen, K. Li, Z. Chen, J. Xia, H. Li, Confined active species and A. Najafpoor, Adsorption and visible-light photocatalytic degradation of
effective charge separation in Bi4O5I2 ultrathin hollow nanotube with increased tetracycline hydrochloride from aqueous solutions using 3D hierarchical
photocatalytic activity, Appl. Catal. B Environ. 268 (2020), 118403, https://doi. mesoporous BiOI: synthesis and characterization, process optimization, adsorption
org/10.1016/j.apcatb.2019.118403. and degradation modeling, Chem. Eng. Res. Des. 129 (2018) 217–230, https://doi.
[47] L. Yang, L. Huang, Y. Li, C. Wang, J. Liu, J. Liu, L. Huang, Y. Song, Y. Lv, In-situ org/10.1016/j.cherd.2017.11.003.
generation Bi12O15Cl6/BiOI core-shell photocatalyst with efficient LED light- [61] H. Lan, G. Zhang, H. Zhang, H. Liu, R. Liu, J. Qu, Solvothermal synthesis of BiOI
driven degradation and antibacterial properties: factors, degradation pathway and flower-like microspheres for efficient photocatalytic degradation of BPA under
mechanism, J. Alloys Compd. 885 (2021), 160884, https://doi.org/10.1016/j. visible light irradiation, Catal. Commun. 98 (2017) 9–12, https://doi.org/
jallcom.2021.160884. 10.1016/j.catcom.2017.02.028.
[48] M. Arumugam, Y. Yu, H.J. Jung, S. Yeon, H. Lee, J. Theerthagiri, S.J. Lee, M. [62] A.C. Mera, Y. Moreno, J.Y. Pivan, O. Peña, H.D. Mansilla, Solvothermal synthesis of
Y. Choi, Solvent-mediated synthesis of BiOI with a tunable surface structure for BiOI microspheres: effect of the reaction time on the morphology and
effective visible light active photocatalytic removal of Cr(VI) from wastewater, photocatalytic activity, J. Photochem. Photobiol. A Chem. 289 (2014) 7–13,
Environ. Res. 197 (2021), 111080, https://doi.org/10.1016/j. https://doi.org/10.1016/j.jphotochem.2014.05.015.
envres.2021.111080. [63] G. Zhang, A. Su, J. Qu, Y. Xu, Synthesis of BiOI flowerlike hierarchical structures
[49] A. Phuruangrat, A. Maneechote, P. Dumrongrojthanath, N. Ekthammathat, toward photocatalytic reduction of CO2 to CH4, Mater. Res. Bull. 55 (2014) 43–47,
S. Thongtem, T. Thongtem, Effect of pH on visible-light-driven Bi2WO6 https://doi.org/10.1016/j.materresbull.2014.04.012.
nanostructured catalyst synthesized by hydrothermal method, Superlattice. [64] L. Zhang, Z. Ma, H. Xu, R. Xie, Y. Zhong, X. Sui, B. Wang, Z. Mao, Preparation of
Microst. 78 (2015) 106–115, https://doi.org/10.1016/j.spmi.2014.11.038. upconversion Yb3+ doped microspherical BiOI with promoted photocatalytic
[50] W. Yu, N. Ji, N. Tian, L. Bai, H. Ou, H. Huang, BiOI/Bi2O2[BO2(OH)] performance, Solid State Sci. 75 (2018) 45–52, https://doi.org/10.1016/j.
heterojunction with boosted photocatalytic degradation performance for diverse solidstatesciences.2017.11.008.
pollutants under visible light irradiation, Colloids Surfaces A Physicochem. Eng. [65] M. Thommes, K. Kaneko, A.V. Neimark, J.P. Olivier, F. Rodriguez-Reinoso,
Asp. 603 (2020), 125184, https://doi.org/10.1016/j.colsurfa.2020.125184. J. Rouquerol, K.S.W. Sing, Physisorption of gases, with special reference to the
[51] Y.R. Jiang, S.Y. Chou, J.L. Chang, S.T. Huang, H.P. Lin, C.C. Chen, Hydrothermal evaluation of surface area and pore size distribution (IUPAC technical report), Pure
synthesis of bismuth oxybromide-bismuth oxyiodide composites with high visible Appl. Chem. 87 (2015) 1051–1069, https://doi.org/10.1515/pac-2014-1117.
light photocatalytic performance for the degradation of CV and phenol, RSC Adv. 5 [66] D.A. Giannakoudakis, N. Farahmand, D. Łomot, K. Sobczak, T.J. Bandosz, J.
(2015) 30851–30860, https://doi.org/10.1039/c5ra01702e. C. Colmenares, Ultrasound-activated TiO2/GO-based bifunctional photoreactive
[52] Y.R. Jiang, H.P. Lin, W.H. Chung, Y.M. Dai, W.Y. Lin, C.C. Chen, Controlled adsorbents for detoxification of chemical warfare agent surrogate vapors, Chem.
hydrothermal synthesis of BiOxCly/BiOmIn composites exhibiting visible-light Eng. J. 395 (2020), 125099, https://doi.org/10.1016/j.cej.2020.125099.
photocatalytic degradation of crystal violet, J. Hazard. Mater. 283 (2015) [67] J. Landers, G.Y. Gor, A.V. Neimark, Density functional theory methods for
787–805, https://doi.org/10.1016/j.jhazmat.2014.10.025. characterization of porous materials, Colloids Surfaces A Physicochem. Eng. Asp.
[53] H. Yu, Q. Han, Effect of reaction mediums on photocatalytic performance of BiOX 437 (2013) 3–32, https://doi.org/10.1016/j.colsurfa.2013.01.007.
(X = Cl, Br, I), Opt. Mater. (Amst). 119 (2021), 111399, https://doi.org/10.1016/j. [68] D.A. Giannakoudakis, A. Qayyum, M. Barczak, R.F. Colmenares-Quintero,
optmat.2021.111399. P. Borowski, K. Triantafyllidis, J.C. Colmenares, Appl. Catal. B Environ. 320
[54] C.W. Siao, W.L.W. Lee, Y.M. Dai, W.H. Chung, J.T. Hung, P.H. Huang, W.Y. Lin, C. (2023), 121939, https://doi.org/10.1016/j.apcatb.2022.121939.
C. Chen, BiO x cl y /BiO m Br n /BiO p I q /GO quaternary composites: syntheses

12

You might also like