ArticlePDF Available

A Novel Low-Cost Photocatalyst: Preparation, Characterization, and Photocatalytic Properties of CeO2-Diatomite Composites

Authors:

Abstract and Figures

Developing CeO2-diatomite composites with highly efficient photocatalytic performance is a practical and low-cost strategy for the removal of abundant contaminants in water and wastewaters. Diatomite (D) was modified by acid treatment to obtain a more porous structure. CeO2-diatomite composites were prepared in two different mass ratios of D to CeO2 of 1:1 and 1:2 via a facile precipitation method. The changes in structural, morphological, optical, and thermal properties of CeO2-diatomite composites were characterized by FTIR, XRD, ESEM-EDAX, BET surface area, TGA, PL, Raman spectroscopy, and zeta potential techniques. ESEM images presented the morphological differences of CeO2-diatomite composites, reflecting the effect of modification as a more folded sheet form morphology and higher BET surface area. XRD analysis revealed the fluorite-type structure of CeO2 particles in composites. Photocatalytic activities were investigated by following the degradation of methylene blue (MB) as a thiazine dye model under UVA light irradiation. CeO2-diatomite composites exhibited irradiation time- and dose-dependent remarkable photocatalytic efficiencies, whereas composite type inconsistent variations were also noticed. The attained performance of the CeO2-diatomite composites could strongly imply a favorable application prospect in the photocatalysis field.
Content may be subject to copyright.
Water 2022, 14, 3373. https://doi.org/10.3390/w14213373 www.mdpi.com/journal/water
Article
A Novel Low-Cost Photocatalyst: Preparation, Characterization,
and Photocatalytic Properties of CeO2-Diatomite Composites
Nazli Turkten
Department of Chemistry, Faculty of Arts and Sciences, Kirsehir Ahi Evran University,
Kirsehir 40100, Turkey; nazli.turkten@ahievran.edu.tr
Abstract: Developing CeO2-diatomite composites with highly efficient photocatalytic performance
is a practical and low-cost strategy for the removal of abundant contaminants in water and
wastewaters. Diatomite (D) was modified by acid treatment to obtain a more porous structure.
CeO2-diatomite composites were prepared in two different mass ratios of D to CeO2 of 1:1 and 1:2
via a facile precipitation method. The changes in structural, morphological, optical, and thermal
properties of CeO2-diatomite composites were characterized by FTIR, XRD, ESEM-EDAX, BET
surface area, TGA, PL, Raman spectroscopy, and zeta potential techniques. ESEM images present-
ed the morphological differences of CeO2-diatomite composites, reflecting the effect of modifica-
tion as a more folded sheet form morphology and higher BET surface area. XRD analysis revealed
the fluorite-type structure of CeO2 particles in composites. Photocatalytic activities were investi-
gated by following the degradation of methylene blue (MB) as a thiazine dye model under UVA
light irradiation. CeO2-diatomite composites exhibited irradiation time- and dose-dependent re-
markable photocatalytic efficiencies, whereas composite type inconsistent variations were also no-
ticed. The attained performance of the CeO2-diatomite composites could strongly imply a favora-
ble application prospect in the photocatalysis field.
Keywords: CeO2-diatomite composites; diatomite; methylene blue; precipitation method;
photocatalysis
1. Introduction
The lack of efficient wastewater treatment and the spreading of hazardous chemi-
cals including dyes from industries into water resources is a major global issue [1]. The
textile industry consumes the most dye, followed by the various dye-related industries,
such as printing, and the production of paper, paint, and leather for various manufac-
turing purposes. Among the different textile dyes, MB is widely used as a cationic and
primary thiazine dye for coloring silk, wool, cotton, and paper [2,3]. The accumulation
of MB in receiving waters can have a harmful impact on the ecological system. There-
fore, the removal of MB from wastewater effluent is highly crucial prior to the direct
discharging into the water bodies [1,3,4]. To date, numerous approaches have been re-
ported for the elimination of MB from wastewaters, including coagulation, flocculation,
biodegradation, and adsorption treatments [3,57]. However, these wastewater treat-
ments are generally not able to achieve complete dye degradation, even with a sequen-
tial combination of two or three methods. Therefore, it is necessary to focus on an alter-
native and efficient water treatment technique having a low cost [811]. Advanced oxi-
dation processes, specifically heterogeneous photocatalysis, may degrade complex or-
ganic pollutants via reactive oxygen species (ROS) operating mainly through hydroxyl
radical attack [12,13].
Recently, new catalysts using either semiconductors such as CdS [14], CuO [15],
MgO [16], SnO2 [17], and Al-doped CoFe2O4 [18], or composites such as PANI-ZnO [19]
Citation: Turkten, N. A Novel
Low-Cost Photocatalyst:
Preparation, Characterization, and
Photocatalytic Properties of
CeO2-Diatomite Composites. Water
2022, 14, 3373. https://doi.org/
10.3390/w14213373
Academic Editor: Chengyun Zhou
Received: 26 September 2022
Accepted: 19 October 2022
Published: 24 October 2022
Publisher’s Note: MDPI stays
neutral with regard to jurisdictional
claims in published maps and
institutional affiliations.
Copyright: © 2022 by the author.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license
(https://creativecommons.org/license
s/by/4.0/).
Water 2022, 14, 3373 2 of 24
and Ag@TiO2/W3 [20], were developed for the photocatalytic degradation of MB. A rare
earth oxide, CeO2, received attention as a promising catalyst since it is non-toxic, non-
photocorrosive, and low-cost, and has a strong oxygen ability with high chemical stabil-
ity properties [2123].
The primary events of a photocatalytic degradation process using a photocatalyst,
e.g., CeO2, can be expressed by the reactions given below (Reactions (1-6)):
Charge separation: Upon irradiation of CeO2, a hole in the valence band (VB, h+VB)
and an electron in the conduction band (CB, eCB) are generated (Reaction (1)).
CeO2 + hν (E > Ebg) CeO2 (eCB + h + VB)
(1)
These charge carriers react with either the H2O molecule (Reaction (2)) or with the
hydroxyl ion OH (Reaction (3)) to form hydroxyl radical (OH).
OH formation via h + VB:
h + VB + H2O H+ + OH
(2)
h + VB + OH OH
(3)
Reduction via eCB:
eCB + O2 O2
(4)
Reaction (4) defines a reduction process indicating formation of a superoxide anion
(O2) followed by formation of H2O2 (Reactions (5) and (6)).
O2 + H2O2 OH + O2 + HO
(5)
OH + OH H2O2
(6)
Degradation of organic pollutants can be achieved through action of ROS (i.e., OH,
O2, H2O2, etc.) proceeding through various consecutive steps and eventually resulting
in formation of innocuous products as CO2, H2O, and inorganic ions (Reaction (7))
[23,24].
ROS + organic pollutants CO2 + H2O
(7)
Recently, some researchers reported the photodegradation of MB using CeO2 as a
catalyst to solve the problem of water pollution [2530]. Considerable efforts were made
to enhance CeO2 adsorption with the addition of SiO2 nanoparticles [23,3135]. Modifica-
tion of the morphology with SiO2 was an advantageous step for successful separation of
the photogenerated charge carriers. It is well-known that the use of SiO2 emerged as a
promising strategy to obtain a more uniform distribution and porosity of the final prod-
uct [31,35].
In the past, D was used to prepare new natural silica mineral carrier materials for
wastewater treatment [3638]. Raw D is a unique ordered porous silica material with a
large specific surface area, low density, high adsorption capacity, and chemical re-
sistance. This low-cost porous material is derived from the skeleton fossils of unicellular
aquatic algae known as diatoms having various shapes and sizes and consisting of
amorphous silica [36,39]. Raw D consists of mineral impurities that reduce the original
structural characteristics. Therefore, the removal of the existing undesired impurities is
sometimes crucial to enhance the pore-size distribution and adsorption properties of D.
Accordingly, acid treatment is applied as a simple and cheap approach, resulting in a
modified diatomite (MD) [3941]. Several studies have focused on the photocatalytic
performance of MB using various catalysts including TiO2-diatomite composites [42,43],
Zn-diatomite composite [44], SnO2-diatomite composite [39], CdS-diatomite composite
[45], and TiO2-ZnO-diatomite composite [46]. However, there were very few studies on
the development of CeO2-diatomite composites. Zhou and colleagues prepared a Pt sin-
gle-atom catalyst on CeO2-modified diatomite support for the hydrogenation of phe-
Water 2022, 14, 3373 3 of 24
nylacetylene to styrene [47]. Su and co-workers proved that the synthesized modified
nano-CeO2-filled oily diatomite/polyvinylidene fluoride composites exhibited tribologi-
cal behavior [48]. Therefore, this study addressed the photocatalytic performance of
CeO2-diatomite composite, which represents a significant contribution.
CeO2-diatomite composites were synthesized via a facile precipitation method con-
sidering the photocatalytic activity of CeO2 with the adsorption ability of diatomite as
the representative material of porous silica. The present work dealt with the preparation
of CeO2-diatomite composites using raw D and MD in two different mass ratios to un-
derstand the relationship between the pretreatment method and resulting composites.
Furthermore, the possible morphological, structural, optical, and thermal differences
depending on the use of the diatomite type in the composite were systematically inves-
tigated. The photocatalytic properties of the prepared CeO2-diatomite composites via
two different routes were evaluated by testing MB dye for the first time.
2. Materials and Methods
2.1. Materials
Diatomite (average particle size = 5.2 μm) was used as supplied from Beg Tug Min-
eral, Türkiye. Chemically, D was mainly composed of SiO2 (75.45%), inorganic constitu-
ents such as Al2O3 (7.15%), CaO (3.05%), Fe2O3 (2.35%), MgO (1.45%), and TiO2 (0.40%),
and some organic impurities. Ce(NH4)2(NO3)6, NaOH, and HCl (37%) were purchased
from Sigma-Aldrich and used without further purification. All aqueous solutions were
prepared using distilled water (conductivity 2 × 10−6 S/m at 25 °C).
2.2. Preparation of Modified Diatomite
Acid-modified diatomite was prepared according to the procedure reported by
Jiang and colleagues [39]. A quantity of 10 g diatomite was added to 100 mL of 1 M HCl
to prepare a suspension and placed in a water bath shaker for 2 h. Then, MD was filtered
through a Gooch funnel equipped with a sintered glass disc, followed by extensive
washings with distilled water, dried at 105 °C for 24 h, and finally calcined in a muffle
furnace at 450 °C in air for 1h.
2.3. Preparation of CeO2-Diatomite Composites
CeO2-diatomite composites were prepared via a modified co-precipitation method
[49]. The theoretical diatomite/CeO2 mass ratios were 1:1 and 1:2 of raw D or MD to
CeO2, and these four composites were denoted as DC-11, DC-12 MDC-11, and MDC-12,
respectively. The representative synthesis of DC-11 was given as follows: 1.72 g raw D
was added to 100 mL of 0.1 M Ce(NH4)2(NO3)6 solution in a flat-bottomed flask and
stirred vigorously for 5 min. Subsequently, the pH was adjusted to 12 by dropwise (ap-
proximately one drop/second) addition of 0.5 M NaOH with vigorous stirring by a mag-
netic stirrer, and stirring was continued for 2 h at room temperature. The precipitate was
filtered, washed with distilled water, and then dried in an air oven at 80 °C for 24 h. Fi-
nally, the dried CeO2-diatomite composites were calcined in a muffle furnace at 500 °C
in air for 2 h. The same synthesis procedure was repeated by taking half the amount of D
for the preparation of DC-12. MDC-11 and MDC-12 composites were synthesized using
MD, accordingly.
2.4. Characterization Techniques
Fourier transform infrared (FTIR) spectroscopy was performed with the attenuated
total reflectance (ATR) technique using a Perkin Elmer Spectrum Two model FTIR and
Universal ATR accessory with diamond/ZnSe crystal in a spectral region of 3600500
cm−1 with a scan resolution of 2 cm−1. Dispersive Raman spectroscopy measurements
were performed by a Thermo Scientific DXR Raman Microscope using Ar+ laser power
(10 mW) at λ = 532 nm. X-ray diffraction (XRD) patterns were obtained by a Rigaku-
Water 2022, 14, 3373 4 of 24
D/MAX-Ultima diffractometer with Cu Kα radiation (λ = 1.54 Å) as an X-ray source. The
accelerating and applied currents were 40 kV and 40 mA, respectively. Scanning electron
microscopy in combination with energy dispersive X-ray analysis (ESEM-EDAX) was
achieved on a FEI-Philips XL30 Scanning Electron Microscope-EDAX instrument.
BrunauerEmmettTeller (BET) surface area and BarrettJoynerHalenda (BJH) pore-
size distribution measurements were obtained by N2 adsorption at 77 K using a
Quantachrome Quadrosorb SI instrument. Thermogravimetric (TG) analysis was under-
taken by a Perkin Elmer model STA 600 in the temperature range of 30 to 800 °C under
pure nitrogen using 10 mg of specimen. The applied heating rate and nitrogen flow rate
were 10 °C/min and 20 mL/min, respectively. The photoluminescence (PL) spectra were
recorded on an Edinburgh Instruments FS5 spectrofluorometer with an excitation wave-
length at λ = 325 nm. Zeta potential measurements were performed in a Malvern
Zetasizer Nano ZSP in the pH range of 212 by adding HC1 or NaOH solutions.
2.5. Assessment of Initial Adsorptive Surface Interactions
Photocatalytic time-dependent (0480 min) adsorption experiments were carried
out in a continuously shaking water bath at 25 ± 1 °C. The volume of MB (5 mg/L) solu-
tion in Erlenmeyer flasks was 25 mL and the adsorbent dose was 0.25 g/L. Following
predetermined time intervals, the suspensions were filtered through 0.45 μm cellulose
acetate membrane filter (Millipore HA) to obtain clear solutions for spectroscopic analy-
sis.
2.6. Photocatalytic Activity Assessment
Photocatalytic activity experiments were carried out in a cylindrical Pyrex reaction
vessel containing 50 mL MB solution illuminated from the top with a 125 W black light
fluorescent lamp (λemis = 300400 nm, λmax = 365 nm). The intensity of the BLF lamp was
determined as Io = 1.65 × 1016 quanta/s using potassium ferrioxalate actinometer [50]. Two
different dosages of CeO2-diatomite composites (0.25 g/L and 0.50 g/L) were used while
keeping the MB concentration constant (5 mg/L). Prior to testing of photocatalytic activi-
ty, the solution was magnetically stirred thoroughly for 30 min in the dark to achieve ef-
fective surface coverage. Following each photocatalytic treatment period (0360 nm), the
solutions were immediately filtrated through 0.45 μm membrane filters. The absorption
changes of MB were assessed at specific maximum absorbance wavelengths of λ = 664
nm, λ = 292 nm, and λ = 246 nm. The UV-vis measurements acquired by a Thermo Scien-
tific Genesys 10S double-beam UV-vis spectrophotometer using 1 cm quartz cells were
denoted as A664, A292, and A246 parameters.
3. Results and Discussion
3.1. Characterization of the Diatomites and CeO2-Diatomite Composites
3.1.1. FTIR Analysis
FTIR spectral features revealed the influence of acid modification on characteristic
functional groups of D, MD, and CeO2-diatomite composites (Figure 1a,b). The spectrum
of D indicated a broad band near 3385 cm−1 that was assigned to the free silanol group
(νSiO–H) on the surface [51]. The bands at 2987 and 2903 cm−1 corresponded to the
asymmetric and symmetric CH bands on CH2 and CH3, respectively, reflecting the
presence of organic matter in diatomite [52,53]. The intense band observed at 1636 cm−1
was related to the bending vibration of water (δHOH). Moreover, the characteristic silica
bands at 1045 and 796 cm−1 were attributed to the asymmetric stretching of SiO–Si as-
SiOSi), and the symmetric stretching vibration of Si–O (νsSiO) in amorphous silica,
respectively [53]. The band at 523 cm−1 was related to the stretching vibration of SiO
due to the deformation on the structural unit of SiO2 [54]. The FTIR spectrum of MD
showed a red shift related to the bands belonging to νasSiO–Si and νsSiO groups of sil-
ica. This could indicate that the acid treatment had an impact on the chemical properties
Water 2022, 14, 3373 5 of 24
of the D. The reason for this could be the change in the chemical composition of MD, re-
sulting in an enhancement of the silica amount due to the removal of the impurity con-
tent of D.
3600 3200 2800 2400 2000 1600 1250 1000 750 500
3385
2903
2987 1636
3386
2989
2901 1629
MD
D
523
796
1045
1051
797
528
Transmittance, a.u.
Wavenumbers, cm1
(a)
3600 3200 2800 2400 2000 1600 1250 1000 750 500
2900
1631
3479
3463
1630
1061
Transmittance, a.u.
Wavenumbers, cm1
3412 2985
2873
2946
2883
2945
2884
(b)
MDC-12
MDC-11
DC-12
DC-11
3445
2989 1055
795
1059
796
1630
798
799
1058
1638
1355
1368
1360
1380
Figure 1. FTIR spectra of (a) D, MD, (b) DC-11, DC-12, MDC-11, and MDC-12.
Water 2022, 14, 3373 6 of 24
The FTIR spectral profiles of composites are shown in Figure 1b. The weak band lo-
cated in the range of 13551380 cm−1 belonged to CeOCe vibration bond in composites,
indicating the existence of CeO2 [55,56]. In general, the positions of the prominent D
bands were slightly shifted in the spectrum of CeO2-diatomite composites. The observed
band shifts could be related to the incorporation of CeO2 into the framework of the SiO2
matrix [31,57]. In the presence of CeO2, a new band related to the stretching vibration of
CeO–Ce (νCe–OCe) at 1380 cm−1 was observed in the spectrum of DC-11 composite.
The main bands at 3445, 2989, 2900, 1631, 1380, 1055, and 795 cm−1 corresponded to the
free silanol group (νSiO–H), asymmetric CH bands on CH2 asCH), symmetric CH
bands on CH3 sCH), bending vibration of water (δHOH), asymmetric stretching of Si
O–Si (νasSiOSi), and symmetric stretching vibration of Si–O (νsSiO), respectively. The
characteristic Si-O-Si asymmetric and symmetric strong stretching signals of D and MD
red shifted in all composites with the increasing amount of CeO2. Although the band at
530 cm−1 was also expected to shift to higher frequencies, it seems to be shielded by the
shoulder of strong O-Ce-O vibrational signal centered at around 415 cm−1 (not shown in
the figure) as reported in the literature [5156]. Moreover, the most interesting spectral
change was observed in the stretching vibration band of CeOCe. This band was nota-
bly blue shifted in MDC-11 and MDC-12 composites compared to that of composites
prepared by raw D, designated as DC-11 and DC-12, respectively.
3.1.2. Raman Spectroscopy
Raman spectra of diatomite specimens and CeO2-diatomite composites are present-
ed in Figure 2.
1000 900 800 700 600 500 400 300 200 100
D
MD
415
416
Intensity, a.u.
Raman Shift, cm1
(a)
Water 2022, 14, 3373 7 of 24
800 700 600 500 400 300 200 100
Intensity, a.u.
Raman Shift, cm1
DC-11
DC-12
MDC-11
MDC-12
464
461
464
466
597
606
596
416
422
418
498
496
533
533
534
(b)
Figure 2. Raman spectra of (a) D, MD, (b) DC-11, DC-12, MDC-11, and MDC-12.
Diatomite specimens exhibited a broad band at ~415 cm−1 that was assigned to the
O-Si-O bending mode [58]. Raman spectra of DC-11 composite revealed a prominent
band at 464 cm−1 that was attributed to the triply degenerate F2g vibration mode, which
was assigned to a symmetric breathing mode of the oxygen atoms around Ce ions [59].
This band was in the range of 461464 cm−1 in composites, suggesting the fluorite struc-
ture of CeO2 [60]. The observed band at ~600 cm−1 corresponded to the non-degenerate
longitudinal-optical mode of CeO2 indicating the oxygen vacancies in the CeO2 lattice
[59,61]. The D1 bands in the region between 500 and 600cm−1 could be related to the bulk
oxygen vacancies in CeO2. The band at ~420 cm−1 could belong to the silica structure due
to the contribution of D. The bands below 200cm−1 could be assigned to the scattering
[62,63].
3.1.3. XRD Analysis
XRD analysis was performed to determine the main chemical composition of raw D
and MD (Figure 3a). XRD patterns of D exhibited a broad diffraction in the region of 2θ
= 2030°, suggesting the presence of amorphous silica. There was no remarkable altera-
tion in the peak positions of the MD. However, a slight enhancement was observed in
the peak intensities, indicating that the acid treatment affected the crystallinity of diato-
mite.
Water 2022, 14, 3373 8 of 24
10 15 20 25 30 35 40 45 50
Intensity, a.u.
2q, Degree
D
MD
K
K
(a)
10 20 30 40 50 60 70 80
DC-11
DC-12
MDC-11
MDC-12
111
•200
220
311
222
400
331
¨K
Intensity, a.u.
2q, Degree
111
•200
220
311
400
331
111
•200
220
311
400
331
111
•200
220
311
400
331
¨K
¨K¨K
(b)
Water 2022, 14, 3373 9 of 24
25 26 27 28 29 30 31 32 33
DC-11
DC-12
MDC-11
MDC-12
28.680
28.800
28.540
28.820
Intensity, a.u.
2q, Degree
(c)
Figure 3. XRD spectra of (a) D, MD, (b) DC-11, DC-12, MDC-11, and MDC-12, (c) enlarged (111)
plane of CeO2.
The XRD patterns of CeO2-diatomite composites are shown Figure 3b. The specific
diffraction peaks were attributed to CeO2 and D. For MDC-12 composite, = 28.68°,
33.19°, 47.85°, 56.72°, 59.42°, 69.69°, and 77.12° corresponded to the presence of (111),
(200), (220), (311), (222), (400), and (331) planes of CeO2, while the broad diffraction peak
at = 13.23 could be related to the kaolinite-like structure in D. The XRD analysis re-
vealed that CeO2 in composites was consistent with the fluorite-type structure due to the
reflection plane of (111) observed at 2θ = 28.68° [64]. The pattern was well-matched with
the CeO2 standard card JCPDS NO. 431002, as well as with Raman data. The diffraction
plane of (222) was not visible for the other three CeO2-diatomite composites. Moreover,
there were no extra diffraction peaks related to the formation of new or mixed phases
between CeO2 and D. The enlarged (111) diffraction planes of CeO2-diatomite compo-
sites were shown in Figure 3c. With increasing amount of D in composites (MDC-11 and
MD-11), a gradual shift towards to higher angles was noted. Compared with DC-12,
the diffraction peak of MDC-12 was also shifted to a higher angle due to acid treatment
process of D.
The average crystallite sizes (D, nm) of CeO2-diatomite composites were calculated
from the peak of the (111) plane of CeO2 using the Scherrer equation (Equation (8)) [65]:
D = K λ/(β cosθ)
(8)
where K = 0.9, λ is the X-ray wavelength (1.5418 Å), θ is the Bragg angle, and β is the full
width at half maximum intensity (FWHM, radians).
The influence of acid treatment on the crystallite size of CeO2-diatomite composites
was also reflected as X-ray line broadening. The diffraction peaks of MDC-12 composite
were obviously narrower than those of DC-12 composite, thus indicating a relatively
larger crystallite size. Almost equal crystal sizes were calculated as 2.2 and 2.1 nm for
DC-11 and MDC-11 composites, respectively. However, the crystallite sizes of CeO2 in
DC-12, of 1.9 nm, and MDC-12, of 3.0 nm, displayed an incremental increase, most prob-
ably due to the modification of diatomite. This result indicates that the acid treatment af-
fected the crystallite particle sizes of CeO2-diatomite composites as the CeO2 amount
was increased.
Water 2022, 14, 3373 10 of 24
3.1.4. ESEM Analysis
The ESEM images of D and MD in comparison to CeO2-diatomite composites are
presented in Figure 4. An orderly set pore structure in a fold sheet form can be observed
in the D image (Figure 4a). The diatom debris grains were arranged dispersedly, result-
ing in a porous structure [66]. After acid treatment, the morphology was altered, and a
more folded sheet form was formed. Moreover, the size of the diatom pores that were
clearly visible was increased (Figure 4b). ESEM images of CeO2-diatomite composites
revealed an irregularly shaped morphology containing both D and CeO2 particles (Fig-
ure 4cf). The dominant morphology was the porous structure of D for DC-11 composite
spreading throughout the composite. It was also clear that a distinctive almost spherical
shape of CeO2 particles could be observed [31]. Increasing the CeO2 amount resulted in a
decrease in the folded sheets of D in DC-12 composite. With respect to ESEM images of
DC-11 and MDC-11 composites, a decrease in both agglomeration and particle size of
CeO2 nanoparticles was evident in MDC-11. The pores of D were enlarged, and folded
sheets almost disappeared. Similar morphological changes were also observed for DC-12
and MDC-12 composites and the particle size was reduced in the MDC-12 composite.
The spherical particles of CeO2 were transformed into various polyhedral particles and
MD was deposited onto the particles. The observed variations could be due to the for-
mation of chemical bonds between CeO2 and diatomite with respect to the CeO2 amount
in composites.
Water 2022, 14, 3373 11 of 24
Figure 4. ESEM images of (a) D, (b) MD, (c) DC-11, (d) DC-12, (e) MDC-11, (f) MDC-12.
EDAX spectral elemental composition of D, MD, and corresponding CeO2-
diatomite composites are displayed in Figure S1, in Supplementary Materials (SI Part 1,
Figure S1af). High content of silicium was strongly confirmed by EDAX analyses of D
and MD. Si content increased after modification and other elements such as C, O, Na,
Mg, Al, Ca, and Fe were also observed. The presence of Cl as detected in the EDAX spec-
trum of MD could be due to the acid modification of D with HCl treatment. The pres-
ence of Ca, Si, and Mg elements could be in the form of CaCO3, silicate, and Mg(OH)2
sediments, as already reported in the materials section (Section 2.1) [67]. More im-
portantly, EDAX spectra (Figure S1cf) showed that the as-prepared composites were
mainly composed of Si, Ce, and O elements.
Water 2022, 14, 3373 12 of 24
3.1.5. BET Analysis
BET nitrogen adsorption/desorption isotherms and BJH pore-size distribution
curves of D, MD, and representative CeO2-diatomite composites are shown in SI Part 1,
Figures S2 and S3. Average pore diameter (d), pore volume (Vp), and specific surface ar-
ea (SBET) are summarized in Table 1. The nitrogen adsorption/desorption isotherms of D
and MD exhibited a type II sorption behavior with an H3 hysteresis loop according to
the International Union of Pure and Applied Chemistry (IUPAC) classification [68]. It
was clear that the D and MD revealed the same form of isotherms and hysteresis loop.
The H3 hysteresis loop was principally related with the filling and emptying of the mes-
opores by capillary condensation, suggesting that the pores were mostly narrow and
slit-shaped, and their sizes were not well-proportioned [69,70]. In addition, the adsorp-
tion isotherm curves were significantly increased with the relatively moderate pressure,
signifying a gradual pore diameter enhancement. However, acid treatment resulted in a
slight enhancement in nitrogen adsorption, indicating that the overflow possessed a
higher SBET. It is well-known that acid modification of D leads to an increase in surface
area and adsorption capacity [7175]. The reason for this could be the elimination of mi-
nor impurities responsible of pore blockage via acid treatment [73].
The adsorption-desorption isotherms of CeO2-diatomite composites were classified
as type IV with an H3 hysteresis loop, implying a typical unfilled mesoporous structure.
However, MDC-12 composite corresponded to a typical IV adsorption isotherm with an
H2 hysteresis loop due to the existence of the pore blockage or seepage [68] (SI Part 1,
Figure S2).
Table 1. Surface characteristics properties of diatomites and CeO2-diatomite composites.
Specimens
SBET *, m2/g
Vp **, cm3/g
D ***, nm
D
109
0.3082
4.55
MD
118
0.3478
5.05
DC-11
76
0.0954
2.29
DC-12
84
0.1578
1.73
MDC-11
127
0.2032
3.83
MDC-12
126
0.1792
3.49
* Specific surface area data calculated by the multi-point BET method. ** Pore volume obtained
from the BJH adsorption cumulative volume of pores between 0.36 and 300 nm diameter. *** Ad-
sorption average pore diameter (4 V/A by BET).
The pore-size distribution curves based on the pore volumes of the BJH adsorption
are shown in SI Part 1, Figure S3. A heterogeneous distribution of pore diameters small-
er than 40 nm was evidenced for D. The DC-11, DC-12, and MDC-11 composites exhibit-
ed narrow pore-size distributions with average pore diameters of 3.82, 2.29, and 1.73 nm,
respectively.
3.1.6. Thermal Analysis
The thermal stability of D and MD and respective DC-11 and MDC-11 composites
were determined by TG and DTG analysis (Figure 5a,b). D and MD exhibited three
weight-loss steps. The first step of TG curves in the low temperature range of 30300 °C
was attributed to the loss of adsorbed water [76]. The weight loss obtained in this step
was 8.5% and 5.4% for D and MD, respectively. The observed peaks at 79 °C (D) and 58
°C (MD) in DTG curves indicated the dehydration process. The second weight loss of D
in the temperature range of 350530 °C with DTG peak was at 458 °C. This MD peak cor-
responding to condensation of hydroxyl groups, and the dehydroxylation process was
slightly shifted to 453 °C [76,77]. The final step was related to the dehydroxylation due
to the loss of OH groups surrounding the Al(VI) atom [78]. A similar DTG peak at ~600
°C was observed for M, MD, and corresponding composites (Figure 5b). The dehydra-
tion of adsorbed H2O and dehydroxylation were also observed in TG curves of compo-
Water 2022, 14, 3373 13 of 24
sites. The total weight loss was 13.2%, 8.02%, 5.31%, and 7.94% for D, MD, DC-11, and
MDC-11, respectively.
3.1.7. PL Analysis
Diatomites were composed of surfaces rich in reactive silanol (SiOH) groups [79
81]. Figure 6 reveals PL spectra of D, MD, and CeO2-composites. The PL spectrum of D
exhibited three main peaks at 400 nm (3.01 eV), 485 nm (2.56 eV), and 564 nm (2.20 eV),
while three peaks belonging to MD were located at 398 nm (3.12 eV), 482 nm (2.57 eV),
and 560 nm (2.21 eV) [82]. It was demonstrated that the origin of visible emission in the
blue and green regions was related to the different defect states, such as oxygen defect
centers, indicating non-bridging oxygen hole centers or neutral oxygen vacancy and self-
trapped excitons [80]. The broad peaks appeared at three main regions of 365435 nm,
450505 nm, and 520600 nm in the PL spectra of composites with different CeO2
amounts. The observed peaks in PL spectra could belong to the characteristic peaks of
CeO2 or D that were also observed in CeO2-diatomite composites with a blue shift. DC-
11 composite showed a strong violet/blue light emission peak at 387 nm (3.20 eV), which
corresponded to the band-edge free exciton luminescence. The blue-green region of visi-
ble range at 467 nm (2.66 eV) was the bound exciton luminescence. The third emission
peak was observed at 564 nm (2.20 eV) in the green visible region [83,84]. The emission
peaks in the range of 400500nm corresponded to the different defect levels between the
localized Ce 4f0 state and O 2p valence band [83,85]. PL emission is useful to understand
the recombination possibility of excited electrons and holes in photocatalyst. In general,
a lower PL intensity indicates a lower recombination rate of electron and hole pairs. Fur-
thermore, an intense PL intensity reveals a higher number of defects and/or oxygen va-
cancies. The observed PL intensity trend with the CeO2 amount in CeO2-diatomite com-
posites was nonlinear. The reason for this could be particle size, recombination, and
presence of defects [23,70,84].
0100 200 300 400 500 600 700 800
86
88
90
92
94
96
98
100
TG,%
Temperature, oC
D
DM
DC-11
MDC-11
(a)
Water 2022, 14, 3373 14 of 24
0100 200 300 400 500 600 700 800
DTG, %/min
D
MD
DC-11
MDC-11
Temperature, oC
622
458
79
58
58
612
453
77
640
638
(b)
Figure 5. (a) TG, (b) DTG curves for D, DM, DC-11, and MDC-11.
360 400 440 480 520 560 600
PL Intensity, a.u.
Wavelength, nm
D
MD
DC-11
DC-12
MDC-11
MDC-12
400
485
564
387
388
387
385
398
Figure 6. PL spectra of D, MD, DC-11, DC-12, MDC-11, and MDC-12.
Water 2022, 14, 3373 15 of 24
3.1.8. Zeta Potential Analysis
pH-dependent variations in zeta potential (ζ, mV), as an efficient tool denoting the
surface charge, are shown in SI Part 1, Figure S4. It was obvious that the zeta potentials
of D and MD had negative magnitudes in the pH range of pH = 212. The surface of MD
was remarkably more negatively charged throughout the pH range compared to D, in-
dicating the presence of a higher number of functional groups. The zeta potentials were
dominantly negative in a wide pH range, excluding the presence of any isoelectric po-
tential (IEP) representing pH of zero-point charge (pHpzc) [86]. This result implies the in-
dependent pH behavior of D and MD particles. Moreover, the monitored permanent
negative charge could be related to the existence of substituted isomorphous alumino-
silicate structure, and this observation was consistent with the chemical/mineralogical
composition of D [87,88]. Similar results were also reported indicating the absence of IEP
[76,89]. In contrast, in several studies, IEP of diatomite was measured as pHpzc = 2 [90,91].
The reason for this could be related to the structure of the diatomite types that contained
different minerals, with various ratios correlating with the zeta potential analyzed. The
pHpzc values of DC-11, DC-12, MDC-11, and MDC-12 composites were 6.07, 4.57, 5.38,
and 5.28, respectively. The results indicate that no remarkable effect of different CeO2
contents was noticed on the surface charges of the MDC-11 and MDC-12 composites.
The reason for this could be the favorable attraction between the positive charge of CeO2
(pHpzc = ~7) and negative charge of MD [23].
3.2. Assessment of Photocatalytic Activity
3.2.1. Preliminary Experiments
The UV-vis absorption spectra of MB revealed four absorption peaks. An intense
absorption peak at λmax = 664 nm (A664) assigned to MB monomer with a shoulder peak at
λ = 612 nm attributed to MB dimer were composed of a sulfurnitrogen conjugated sys-
tem expressing the chromophore groups of MB. The two other observed peaks at λ = 292
nm (A292) and λ = 246 nm (A246) were related to the substituted benzene rings, namely,
the phenothiazine unit in MB structure [3]. Photocatalytic degradation of MB was ex-
pressed by descriptive parameters such as A664, A292, and A246.
The preliminary experiments displayed direct photolysis of MB and surface interac-
tions, revealing MB adsorption onto CeO2-diatomite composites. As a representative pa-
rameter of MB, A664 was used unless otherwise stated [92].
Variations in UV-vis spectral features of MB under irradiation due to direct photo-
degradation were formerly reported by Turkten and colleagues [19]. Accordingly, direct
photodegradation of MB was extremely slow due to the non-overlapping trend of UV-
vis absorption spectra with the BLF lamp emission spectra.
The effect of pre-contact time prior to initiation of light exposure was determined
under dark conditions. D and DC-11 composite (0.25 g/L) were chosen as representative
basic specimens and adsorption of MB (5 mg/L) was investigated for a range of contact
time (t = 08 h). The initial adsorption amount was determined as ~20%, as related to the
prevailing interactions between the negatively charged surface of the specimens and cat-
ionic dye MB [93]. Slight variations 10%) were attained during the whole period of
dark interactions. Fundamentally, the extent of initial adsorption should sufficiently dis-
close the free surface area for light absorption required for the initiation of the photo-
catalytic reaction. Therefore, as an initial dark contact time, 30 min was chosen for pho-
tocatalytic treatment of MB using all specimens.
3.2.2. Photocatalytic Degradation of MB using CeO2-Diatomite Composites
The photocatalytic degradation of MB was studied using CeO2-diatomite compo-
sites using two different doses (0.25 and 0.50 g/L). As an effective constraint of the sur-
face-oriented nature of photocatalysis prior to initiation of light exposure, the dose-
Water 2022, 14, 3373 16 of 24
dependent initial dark adsorption extents of all descriptive parameters are presented in
Figure 7.
Figure 7. Dose dependent initial dark adsorption extents of MB onto DC-11, DC-12, MDC-11, and
MDC-12, as expressed by all descriptive parameters.
From a general perspective, coverage of MB onto composites increased with respect
to increasing dose for the MDC-11 specimen. Upon use of the lower dose of 0.25 g/L, all
descriptive parameters displayed a consistently decreasing trend, irrespective of the
composite type. However, for the higher dose of 0.50 g/L, all descriptive parameters dis-
played similar roles towards the surface of the photocatalyst specimens. The MDC-11
specimen expressed a remarkably different trend as evidenced by the highest surface in-
teraction, most probably due to the counter-balancing effect related to the equal compo-
sitional ratio of MD to CeO2.
Upon initiation of irradiation, during the whole duration of photocatalysis (up to
300 min), UV-vis spectral features of MB were followed, as shown in SI Part II, Figure S5
and Figure S6. The gradual decrement in the A664 peak during photocatalysis implied
that the sulfurnitrogen conjugated system was degraded. A blue shift was noticed due
to the prolonged exposure of UVA illumination. This result indicated a probable route of
N-demethylated degradation of MB derivatives [19,94,95]. A similar progressively de-
creasing trend in A292 and A246 was also recorded. This outcome indicated the occurrence
of oxidative destruction of MB, resulting in a ring-opening reaction of the phenothiazine
species. The absence of new peaks suggested that there was no formation of new inter-
mediates converted from the destroyed phenothiazine structure of MB [94,96]. At 300
min of irradiation, all descriptive parameters were diminished and almost a complete
disappearance of the A612 was found upon use of the MDC-11 composite. This could be
ascribed to the cleavage of benzene rings and the heteropoly aromatic linkage due to the
degradation of MB.
The photocatalytic degradation of MB in the presence of the prepared CeO2-
diatomite composites followed the pseudo-first-order rate model expressed by Equation
(9):
Rate (R) = dA/dt = kA
(9)
where R is the pseudo-first-order rate (cm−1min−1); Ao is the initial absorbance of MB ex-
pressed as A664,o, A292,o and A246,o; A is the absorbance of MB expressed as A664, A292, and
A246 at time t; t is the irradiation time (min); and k is the pseudo-first-order reaction rate
constant (min−1). All UV-vis parameters, i.e., A664, A292, and A246, exhibited logarithmic
decay profiles, revealing kinetic model parameters as presented in Table 2.
Water 2022, 14, 3373 17 of 24
Table 2. Photocatalytic degradation kinetics of MB expressed by A664, A292, and A246.
First-Order Kinetic Parameters
Dose: 0.25 g/L
First-Order Kinetic Parameters
Dose: 0.50 g/L
A664
kx103,
min1
t1/2,
min
Rx103,
cm1 min1
kx102,
min1
t1/2,
min
Rx103,
cm1 min1
DC-11
2.24
309
2.15
5.91
117
5.68
DC-12
3.54
196
3.40
8.98
77
8.63
MDC-11
10.2
68
9.84
9.00
77
8.65
MDC-12
3.05
227
2.93
5.36
129
5.15
A292
DC-11
1.93
359
1.05
5.00
139
2.72
DC-12
2.76
251
1.50
6.56
106
3.57
MDC-11
2.76
251
1.50
6.11
113
3.32
MDC-12
2.49
278
1.35
4.33
160
2.35
A246
DC-11
1.42
488
0.344
6.19
112
1.50
DC-12
1.83
379
0.443
7.41
94
1.80
MDC-11
3.77
184
0.913
8.58
81
2.07
MDC-12
1.64
423
0.397
4.22
164
1.02
The MDC-11 composite expressed the highest decolorization rate constant upon use
of two different doses, and thus the highest photocatalytic activity. The reason for this
could be the larger specific surface area of the MDC-11 composite (SBET = 127 m2/g),
providing more available active sites, and hence favorable for the fast adsorption of MB,
resulting in a high photocatalytic activity [97].
The photocatalytic degradation rate constants of CeO2-diatomite composites using a
0.25 g/L photocatalyst dose can be presented in decreasing order as:
A664: MDC-11 > DC-12 > MDC-12 > DC-11
A292: MDC-11 = DC-12 > MDC-12 > DC-11
A246: MDC-11 > DC-12 > MDC-12 > DC-11
Upon use of a 0.50 g/L photocatalyst dose, the trend in photocatalytic degradation
rate constants can be given in the following order as:
A664: MDC-11 > DC-12 > DC-11 > MDC-12
A292: DC-12 > MDC-11 > DC-11 > MDC-12
A246: MDC-11 > DC-12 > DC-11 > MDC-12
The degree of MB decolorization/degradation using CeO2-diatomite composites, as
presented in Figure 8, was also calculated by the represented Equations (1012):
Decolorization664, % = ((A664,o A664)/ A664,o)) × 100
(10)
Degradation292, % = ((A292,o A292)/ A292,o)) × 100
(11)
Degradation246, % = ((A246,o A246)/ A246,o)) × 100
(12)
Water 2022, 14, 3373 18 of 24
Figure 8. Dose-dependent decolorization/degradation efficiencies of MB onto DC-11, DC-12,
MDC-11, and MDC-12 as expressed by all descriptive parameters.
The degree of MB decolorization/degradation using 0.25 g/L of CeO2-diatomite
composites for an irradiation period of 120 min can be presented in a decreasing order
as:
A664: MDC-11 > DC-12 > MDC-12 > DC-11
A292: MDC-11 > DC-12 > MDC-12 > DC-11
A246: MDC-11 > DC-12 > MDC-12 > DC-11
The degree of MB decolorization/degradation using 0.50 g/L of CeO2-diatomite
composites for an irradiation period of 120 min can be presented in the following order
as:
A664: DC-12 > MDC-11 > DC-11 > MDC-12
A292: DC-12 > MDC-11 > DC-11 > MDC-12
A246: MDC-11 > DC-12 > DC-11 > MDC-12
Upon use of 0.25 g/L, the effect of diatomite modification was found to be more
pronounced for DC-11 and MDC-11 specimens. The lowest decolorization/degradation
efficiency attained for DC-11 can be expressed by the folded diatomite structure (Figure
4a) with a high agglomeration resulting in a negative effect on the photocatalytic effi-
ciency [98].
An increased CeO2 ratio resulted in a retardation effect, as observed for DC-12 and
MDC-12; the reason of this could be the lower surface coverage related to the surface
morphological properties (Table 1, Figure 7). However, the presence of a higher photo-
catalyst dose (0.50 g/L) resulted in a pronounced effect due to both diatomite modifica-
tion and CeO2 content, as well as the available surface area (Figure 8).
For simplicity purposes, as an indicative parameter of MB, A664 was selected for
comparative presentation of kinetics and removal of MB upon use of all photocatalyst
specimens (Figure 9).
Water 2022, 14, 3373 19 of 24
Figure 9. Comparison of dose-dependent initial adsorption extents and decolorization efficiencies
MB, A664 with respect to first-order kinetic rate constants.
As can be visualized, upon use of 0.25 g/L, first-order rate constants revealed an in-
consistent relationship with surface coverage extents, whereas decolorization efficiencies
represented a close resemblance irrespective of the photocatalyst type. The highest rate
constant of MDC-11 could not be directly correlated with the surface coverage extent,
although decolorization efficiency was found to be in good order. On the other hand,
upon use of 0.50 g/L, first-order rate constants indicated differing tendencies with re-
spect to initial adsorption extents; however, they displayed similarity to the observed
decolorization efficiencies. Increasing the photocatalyst dose substantially increased
degradation rates, especially in the MDC-11 specimen, which could be related to the
degradation mechanism resulting in co-operating adsorption based on Langmuir
Hinshelwood kinetics and, in solution, operating in the EleyRideal kinetic model.
It should be emphasized that photocatalytic degradation of methylene blue should
be followed by all descriptive spectroscopic parameters rather than sole decolorization
(A664) due to hydroxylation leading to the aromatic ring opening [92].
4. Conclusions
CeO2-composites were prepared by a simple precipitation method using diatomite
and modified diatomite as representative silica sources. The effect of acid treatment re-
sulted in CeO2-composites having different morphological, structural, optical, and ther-
mal features. The application pretreatment of diatomite not only led to the formation of
a less-agglomerated folded sheet structure, but also enhanced the surface area and po-
rosity of MDC-11 composite compared to DC-11 composite. The presence of the FTIR
bands belonging to the CeOCe vibration bond and silica groups confirmed the cou-
pled binary system of CeO2 and SiO2 in composites. The XRD and Raman analysis indi-
cated that CeO2 resembled a fluorite-type structure in the composites. The average crys-
tallite sizes of CeO2 in the composites determined through Scherrer’s equation were in
the range of 1.93.0 nm. The thermal behavior of the composites was related to the de-
hydration of adsorbed water and dihydroxylation.
The photocatalytic activity of the prepared CeO2-diatomite composites was exam-
ined through the degradation of MB, represented by three descriptive absorbances, i.e.,
A664, A292, and A246. The highest photocatalytic efficiency of MB was obtained using the
MDC-11 composite due to the high surface area of 127 m2/g, resulting in enhanced sur-
face contact between MB and the composite.
Water 2022, 14, 3373 20 of 24
This study originally described a systematic comparison effect of acid treatment of
diatomite on the preparation of CeO2-diatomite composites, as well as photocatalytic ac-
tivities. Based on the attained results using MB as the substrate, CeO2-diatomite compo-
sites could be a promising new affordable and low-cost candidate for photocatalytic ap-
plications.
Supplementary Materials: The following supporting information can be downloaded at:
https://www.mdpi.com/article/10.3390/w14213373/s1, Figure S1: EDAX spectra of (a) D, (b) MD,
(c) DC-11, (d) DC-12, (e) MDC-11, (f) MDC-12; Figure S2: BET adsorption-desorption isotherms of
(a) D, (b) MD, (c) DC-11, (d) MDC-11; Figure S3: Pore size distributions of (a) D, (b) MD, (c) DC-11,
(d) MDC-11; Figure S4: Zeta potential plots of (a) diatomites, (b) CeO2-diatomite composites as a
function of pH; Figure S5: UV-vis absorption spectra of MB using 0.25 mg/mL dose of (a) DC-11,
(b) DC-12, (c) MDC-11, and (d) MDC-12; Figure S6: UV-vis absorption spectra of MB using 0.50
mg/mL dose of (a) DC-11, (b) DC-12, (c) MDC-11, and (d) MDC-12.
Funding: This research received no external funding.
Data Availability Statement: The data presented in the study is available on request from the cor-
responding author.
Acknowledgments: The author is thankful to Beg Tug Mineral, Türkiye for generous support.
Conflicts of Interest: The author declares that she has no known competing financial interests or
personal relationships that could have appeared to influence the work reported in this paper.
References
1. Waghchaure, R.H.; Adole, V.A.; Jagdale, B.S. Photocatalytic degradation of methylene blue, rhodamine B, methyl orange and
Eriochrome black T dyes by modified ZnO nanocatalysts: A concise review. Inorg. Chem. Commun. 2022, 143, 109764.
https://doi.org/10.1016/j.inoche.2022.109764.
2. Hamad, H.N.; Idrus, S. Recent developments in the application of bio-waste-derived adsorbents for the removal of methylene
blue from wastewater: A review. Polymers 2022, 14, 783.
3. Khan, I.; Saeed, K.; Zekker, I.; Zhang, B.; Hendi, A.H.; Ahmad, A.; Ahmad, S.; Zada, N.; Ahmad, H.; Shah, L.A.; et al. Review
on methylene blue: Its properties, uses, toxicity and photodegradation. Water 2022, 14, 242.
4. Rafatullah, M.; Sulaiman, O.; Hashim, R.; Ahmad, A. Adsorption of methylene blue on low-cost adsorbents: A review. J.
Hazard. Mater. 2010, 177, 7080. https://doi.org/10.1016/j.jhazmat.2009.12.047.
5. Kuang, Y.; Zhang, X.; Zhou, S. Adsorption of methylene blue in water onto activated carbon by surfactant modification. Water
2020, 12, 587.
6. Labena, A.; Abdelhamid, A.E.; Amin, A.S.; Husien, S.; Hamid, L.; Safwat, G.; Diab, A.; Gobouri, A.A.; Azab, E. Removal of
methylene blue and congo red using adsorptive membrane impregnated with dried Ulva fasciata and Sargassum dentifolium.
Plants 2021, 10, 384.
7. Verma, A.K.; Dash, R.R.; Bhunia, P. A review on chemical coagulation/flocculation technologies for removal of colour from
textile wastewaters. J. Environ. Manage. 2012, 93, 154168. https://doi.org/10.1016/j.jenvman.2011.09.012.
8. Gomes, J.; Lincho, J.; Domingues, E.; Quinta-Ferreira, R.M.; Martins, R.C. NTiO2 Photocatalysts: A Review of Their
Characteristics and Capacity for Emerging Contaminants Removal. Water 2019, 11, 373.
9. Kumar, M.S.; Sonawane, S.H.; Pandit, A.B. Degradation of methylene blue dye in aqueous solution using hydrodynamic
cavitation based hybrid advanced oxidation processes. Chem. Eng. Process. Process Intensif. 2017, 122, 288295.
https://doi.org/10.1016/j.cep.2017.09.009.
10. Kurian, M. Advanced oxidation processes and nanomaterialsA review. Cleaner Eng. Technol. 2021, 2, 100090.
https://doi.org/10.1016/j.clet.2021.100090.
11. Modi, S.; Yadav, V.K.; Gacem, A.; Ali, I.H.; Dave, D.; Khan, S.H.; Yadav, K.K.; Rather, S.-u.; Ahn, Y.; Son, C.T.; et al. Recent
and Emerging Trends in Remediation of Methylene Blue Dye from Wastewater by Using Zinc Oxide Nanoparticles. Water
2022, 14, 1749.
12. Hoffmann, M.R.; Martin, S.T.; Choi, W.; Bahnemann, D.W. Environmental Applications of Semiconductor Photocatalysis.
Chem. Rev. 1995, 95, 6996. https://doi.org/10.1021/cr00033a004.
13. Turchi, C.S.; Ollis, D.F. Photocatalytic degradation of organic water contaminants: Mechanisms involving hydroxyl radical
attack. J. Catal. 1990, 122, 178192. https://doi.org/10.1016/0021-9517(90)90269-P.
14. Gharbani, P.; Mehrizad, A.; Mosavi, S.A. Optimization, kinetics and thermodynamics studies for photocatalytic degradation
of Methylene Blue using cadmium selenide nanoparticles. npJ. Clean Water 2022, 5, 34. https://doi.org/10.1038/s41545-022-
00178-x.
15. George, A.; Magimai Antoni Raj, D.; Venci, X.; Dhayal Raj, A.; Albert Irudayaraj, A.; Josephine, R.L.; John Sundaram, S.; Al-
Mohaimeed, A.M.; Al Farraj, D.A.; Chen, T.-W.; et al. Photocatalytic effect of CuO nanoparticles flower-like 3D nanostructures
Water 2022, 14, 3373 21 of 24
under visible light irradiation with the degradation of methylene blue (MB) dye for environmental application. Environ. Res.
2022, 203, 111880. https://doi.org/10.1016/j.envres.2021.111880.
16. Ahmad, A.; Khan, M.; Khan, S.; Luque, R.; Almutairi, T.M.; Karami, A.M. Bio-construction of MgO nanoparticles using Texas
sage plant extract for catalytical degradation of methylene blue via photocatalysis. Int. J. Environ. Sci. Technol. 2022, 112.
https://doi.org/10.1007/s13762-022-04090-2.
17. Perumal, V.; Inmozhi, C.; Uthrakumar, R.; Robert, R.; Chandrasekar, M.; Mohamed, S.B.; Honey, S.; Raja, A.; Al-Mekhlafi,
F.A.; Kaviyarasu, K. Enhancing the photocatalytic performance of surfaceTreated SnO2 hierarchical nanorods against
methylene blue dye under solar irradiation and biological degradation. Environ. Res. 2022, 209, 112821.
https://doi.org/10.1016/j.envres.2022.112821.
18. Abbas, N.; Rubab, N.; Sadiq, N.; Manzoor, S.; Khan, M.I.; Fernandez Garcia, J.; Barbosa Aragao, I.; Tariq, M.; Akhtar, Z.;
Yasmin, G. Aluminum-Doped Cobalt Ferrite as an Efficient Photocatalyst for the Abatement of Methylene Blue. Water 2020,
12, 2285.
19. Turkten, N.; Karatas, Y.; Bekbolet, M. Preparation of PANI Modified ZnO Composites via Different Methods: Structural,
Morphological and Photocatalytic Properties. Water 2021, 13, 1025. https://doi.org/10.3390/w13081025.
20. Basumatary, B.; Basumatary, R.; Ramchiary, A.; Konwar, D. Evaluation of Ag@TiO2/WO3 heterojunction photocatalyst for
enhanced photocatalytic activity towards methylene blue degradation. Chemosphere 2022, 286, 131848.
https://doi.org/10.1016/j.chemosphere.2021.131848.
21. Birben, N.C.; Paganini, M.C.; Calza, P.; Bekbolet, M. Photocatalytic degradation of humic acid using a novel photocatalyst: Ce-
doped ZnO. Photochem. Photobiol. Sci. 2017, 16, 2430. https://doi.org/10.1039/C6PP00216A.
22. Chen, F.; Cao, Y.; Jia, D. Preparation and photocatalytic property of CeO2 lamellar. Appl. Surf. Sci. 2011, 257, 92269231.
https://doi.org/10.1016/j.apsusc.2011.06.009.
23. Phanichphant, S.; Nakaruk, A.; Channei, D. Photocatalytic activity of the binary composite CeO2/SiO2 for degradation of dye.
Appl. Surf. Sci. 2016, 387, 214220. https://doi.org/10.1016/j.apsusc.2016.06.072.
24. Litter, M.I. Heterogeneous photocatalysis: Transition metal ions in photocatalytic systems. Appl. Catal. B 1999, 23, 89114.
https://doi.org/10.1016/S0926-3373(99)00069-7.
25. Cheng, Z.; Luo, S.; Liu, Z.; Zhang, Y.; Liao, Y.; Guo, M.; Nguyen, T.T. Visible-light-driven hierarchical porous CeO2 derived
from wood for effective photocatalytic degradation of methylene blue. Opt. Mater. 2022, 129, 112429.
https://doi.org/10.1016/j.optmat.2022.112429.
26. Jasim, S.A.; Machek, P.; Abdelbasset, W.K.; Jarosova, M.; Majdi, H.S.; Khalaji, A.D. Solution combustion synthesis of CeO2
nanoparticles for excellent photocatalytic degradation of methylene blue. Appl. Phys. A 2022, 128, 475.
https://doi.org/10.1007/s00339-022-05532-x.
27. Majumder, D.; Chakraborty, I.; Mandal, K.; Roy, S. Facet-Dependent Photodegradation of Methylene Blue Using Pristine CeO2
Nanostructures. ACS Omega 2019, 4, 42434251. https://doi.org/10.1021/acsomega.8b03298.
28. Pouretedal, H.R.; Kadkhodaie, A. Synthetic CeO2 Nanoparticle Catalysis of Methylene Blue Photodegradation: Kinetics and
Mechanism. Chinese J. Catal. 2010, 31, 13281334. https://doi.org/10.1016/S1872-2067(10)60121-0.
29. Reddy Yadav, L.S.; Lingaraju, K.; Daruka Prasad, B.; Kavitha, C.; Banuprakash, G.; Nagaraju, G. Synthesis of CeO2
nanoparticles: Photocatalytic and antibacterial activities. Eur. Phys. J. Plus 2017, 132, 239. https://doi.org/10.1140/epjp/i2017-
11462-4.
30. Saadoon, S.J.; Jarosova, M.; Machek, P.; Kadhim, M.M.; Ali, M.H.; Khalaji, A.D. Methylene blue photodegradation using as-
synthesized CeO2 nanoparticles. J. Chin. Chem. Soc. 2022, 69, 280288. https://doi.org/10.1002/jccs.202100476.
31. Christy, E.J.S.; Alagar, R.; Dhanu, M.; Pius, A. Porous nonhierarchical CeO2/SiO2 monolith for effective degradation of organic
pollutants. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100365. https://doi.org/10.1016/j.enmm.2020.100365.
32. Dhmees, A.S.; Rashad, A.M.; Eliwa, A.A.; Zawrah, M.F. Preparation and characterization of nano SiO2@CeO2 extracted from
blast furnace slag and uranium extraction waste for wastewater treatment. Ceram. Int. 2019, 45, 73097317.
https://doi.org/10.1016/j.ceramint.2019.01.014.
33. Farrukh, M.A.; Butt, K.M.; Chong, K.-K.; Chang, W.S. Photoluminescence emission behavior on the reduced band gap of Fe
doping in CeO2-SiO2 nanocomposite and photophysical properties. J. Saudi Chem. Soc. 2019, 23, 561575.
https://doi.org/10.1016/j.jscs.2018.10.002.
34. Mohamed, R.M.; Aazam, E.S. Synthesis and Characterization of CeO2-SiO2 Nanoparticles by Microwave-Assisted Irradiation
Method for Photocatalytic Oxidation of Methylene Blue Dye. Int. J. Photoenergy 2012, 2012, 928760.
https://doi.org/10.1155/2012/928760.
35. Rani, N.; Ahlawat, R.; Goswami, B. Annealing effect on bandgap energy and photocatalytic properties of CeO2-SiO2
nanocomposite prepared by sol-gel technique. Mater. Chem. Phys. 2020, 241, 122401.
https://doi.org/10.1016/j.matchemphys.2019.122401.
36. Li, X.; Simon, U.; Bekheet, M.F.; Gurlo, A. Mineral-Supported Photocatalysts: A Review of Materials, Mechanisms and
Environmental Applications. Energies 2022, 15, 5607.
37. Mishra, A.; Mehta, A.; Basu, S. Clay supported TiO2 nanoparticles for photocatalytic degradation of environmental pollutants:
A review. J. Environ. Chem. Eng. 2018, 6, 60886107. https://doi.org/10.1016/j.jece.2018.09.029.
Water 2022, 14, 3373 22 of 24
38. Tang, X.; Tang, R.; Xiong, S.; Zheng, J.; Li, L.; Zhou, Z.; Gong, D.; Deng, Y.; Su, L.; Liao, C. Application of natural minerals in
photocatalytic degradation of organic pollutants: A review. Sci. Total Environ. 2022, 812, 152434.
https://doi.org/10.1016/j.scitotenv.2021.152434.
39. Jiang, H.; Wang, R.; Wang, D.; Hong, X.; Yang, S. SnO2/Diatomite Composite Prepared by Solvothermal Reaction for Low-
Cost Photocatalysts. Catalysts 2019, 9, 1060.
40. Xiong, C.; Ren, Q.; Chen, S.; Liu, X.; Jin, Z.; Ding, Y. A multifunctional Ag3PO4/Fe3O4/Diatomite composites: Photocatalysis,
adsorption and sterilization. Mater. Today Commun. 2021, 28, 102695. https://doi.org/10.1016/j.mtcomm.2021.102695.
41. Zhang, G.; Cai, D.; Wang, M.; Zhang, C.; Zhang, J.; Wu, Z. Microstructural modification of diatomite by acid treatment, high-
speed shear, and ultrasound. Microporous Mesoporous Mater. 2013, 165, 106112.
https://doi.org/10.1016/j.micromeso.2012.08.005.
42. Cherrak, R.; Hadjel, M.; Benderdouche, N.; Bellayer, S.; Traisnel, M. Treatment of recalcitrant organic pollutants in water by
heterogeneous catalysis using a mixed material (TiO2-diatomite of algeria). Desalination Water Treat. 2016, 57, 1713917148.
https://doi.org/10.1080/19443994.2016.1162201.
43. Zhang, G.; Wang, B.; Sun, Z.; Zheng, S.; Liu, S. A comparative study of different diatomite-supported TiO2 composites and
their photocatalytic performance for dye degradation. Desalination Water Treat. 2016, 57, 1751217522.
https://doi.org/10.1080/19443994.2015.1085449.
44. Tanniratt, P.; Wasanapiarnpong, T.; Mongkolkachit, C.; Sujaridworakun, P. Utilization of industrial wastes for preparation of
high performance ZnO/diatomite hybrid photocatalyst. Ceram. Int. 2016, 42, 1760517609.
https://doi.org/10.1016/j.ceramint.2016.08.074.
45. Tajmiri, S.; Hosseini, M.R.; Azimi, E. Combined photocatalytic-adsorptive removal of water contaminants using a biologically
prepared CdS-diatomite nanocomposite. Mater. Chem. Phys. 2021, 258, 123913.
https://doi.org/10.1016/j.matchemphys.2020.123913.
46. Yang, B.; Ma, Z.; Wang, Q.; Yang, J. Synthesis and Photoelectrocatalytic Applications of TiO2/ZnO/Diatomite Composites.
Catalysts 2022, 12, 268.
47. Zhou, Y.; Xi, W.; Xie, Z.; You, Z.; Jiang, X.; Han, B.; Lang, R.; Wu, C. High-Loading Pt Single-Atom Catalyst on CeO2-Modified
Diatomite Support. Chem. Asian J. 2021, 16, 26222625. https://doi.org/10.1002/asia.202100730.
48. Su, C.; Wang, H.; Li, M.; Qu, Y.; Zhu, Y. Tribological behavior and characterization analysis of modified nano-CeO2 filled oily
diatomite/PVDF composites. Tribol. Int. 2019, 130, 299307. https://doi.org/10.1016/j.triboint.2018.10.007.
49. Seeharaj, P.; Kongmun, P.; Paiplod, P.; Prakobmit, S.; Sriwong, C.; Kim-Lohsoontorn, P.; Vittayakorn, N. Ultrasonically-
assisted surface modified TiO2/rGO/CeO2 heterojunction photocatalysts for conversion of CO2 to methanol and ethanol.
Ultrason. Sonochem. 2019, 58, 104657. https://doi.org/10.1016/j.ultsonch.2019.104657.
50. Hatchard, C.G.; Parker, C.A.; Bowen, E.J. A new sensitive chemical actinometerII. Potassium ferrioxalate as a standard
chemical actinometer. Proc. R. Soc. Lond. A 1956, 235, 518536. https://doi.org/10.1098/rspa.1956.0102.
51. Dai, X.; Zeng, H.; Jin, C.; Rao, J.; Liu, X.; Li, K.; Zhang, Y.; Yu, Y.; Zhang, Y. 2D3D graphene-coated diatomite as a support
toward growing ZnO for advanced photocatalytic degradation of methylene blue. RSC Adv. 2021, 11, 3850538514.
https://doi.org/10.1039/D1RA07708B.
52. Ma, L.; Xu, H.; Xie, Q.; Chen, N.; Yu, Q.; Li, C. Mechanism of As(V) adsorption from aqueous solution by chitosan-modified
diatomite adsorbent. J. Dispersion Sci. Technol. 2021, 43, 15121524. https://doi.org/10.1080/01932691.2021.1876592.
53. Pu, X.; Dang, Q.; Liu, C.; Xu, Q.; Li, B.; Ji, X.; Liu, H.; Ma, Y.; Zhang, B.; Cha, D. Selective capture of mercury(II) in aqueous
media using nanoporous diatomite modified by allyl thiourea. J. Mater. Sci. 2022, 57, 92469264.
https://doi.org/10.1007/s10853-022-07245-1.
54. Nguyen, Q.-B.; Vahabi, H.; Rios de Anda, A.; Versace, D.-L.; Langlois, V.; Perrot, C.; Nguyen, V.-H.; Naili, S.; Renard, E. Dual
UV-Thermal Curing of Biobased Resorcinol Epoxy Resin-Diatomite Composites with Improved Acoustic Performance and
Attractive Flame Retardancy Behavior. Sustain. Chem. 2021, 2, 2448.
55. Othman, A.; Vargo, P.; Andreescu, S. Recyclable Adsorbents Based on Ceria Nanostructures on Mesoporous Silica Beads for
the Removal and Recovery of Phosphate from Eutrophic Waters. ACS Appl. Nano Mater. 2019, 2, 70087018.
https://doi.org/10.1021/acsanm.9b01512.
56. Xie, L.; Ren, Z.; Zhu, P.; Xu, J.; Luo, D.; Lin, J. A novel CeO2TiO2/PANI/NiFe2O4 magnetic photocatalyst: Preparation,
characterization and photodegradation of tetracycline hydrochloride under visible light. J. Solid State Chem. 2021, 300, 122208.
https://doi.org/10.1016/j.jssc.2021.122208.
57. Laha, S.C.; Mukherjee, P.; Sainkar, S.R.; Kumar, R. Cerium Containing MCM-41-Type Mesoporous Materials and their Acidic
and Redox Catalytic Properties. J. Catal. 2002, 207, 213223. https://doi.org/10.1006/jcat.2002.3516.
58. Yao, G.; Lei, J.; Zhang, X.; Sun, Z.; Zheng, S.; Komarneni, S. Mechanism of zeolite X crystallization from diatomite. Mater. Res.
Bull. 2018, 107, 132138. https://doi.org/10.1016/j.materresbull.2018.07.021.
59. Reddy, B.M.; Khan, A.; Yamada, Y.; Kobayashi, T.; Loridant, S.; Volta, J.-C. Surface Characterization of CeO2/SiO2 and
V2O5/CeO2/SiO2 Catalysts by Raman, XPS, and Other Techniques. J. Phys. Chem. B 2002, 106, 1096410972.
https://doi.org/10.1021/jp021195v.
60. Mamontov, G.V.; Grabchenko, M.V.; Sobolev, V.I.; Zaikovskii, V.I.; Vodyankina, O.V. Ethanol dehydrogenation over Ag-
CeO2/SiO2 catalyst: Role of Ag-CeO2 interface. Appl. Catal. A 2016, 528, 161167. https://doi.org/10.1016/j.apcata.2016.10.005.
Water 2022, 14, 3373 23 of 24
61. Weber, W.H.; Hass, K.C.; McBride, J.R. Raman study of CeO2: Second-order scattering, lattice dynamics, and particle-size
effects. Phys. Rev. B 1993, 48, 178185. https://doi.org/10.1103/PhysRevB.48.178.
62. Grabchenko, M.V.; Mamontov, G.V.; Zaikovskii, V.I.; La Parola, V.; Liotta, L.F.; Vodyankina, O.V. Design of Ag-CeO2/SiO2
catalyst for oxidative dehydrogenation of ethanol: Control of Ag-CeO2 interfacial interaction. Catal. Today 2019, 333, 29.
https://doi.org/10.1016/j.cattod.2018.05.014.
63. Humbert, B.; Burneau, A.; Gallas, J.P.; Lavalley, J.C. Origin of the Raman bands, D1 and D2, in high surface area and vitreous
silicas. J. Non-Cryst. Solids 1992, 143, 7583. https://doi.org/10.1016/S0022-3093(05)80555-1.
64. Cerrato, E.; Calza, P.; Cristina Paganini, M. Photocatalytic reductive and oxidative ability study of pristine ZnO and CeO2-
ZnO heterojunction impregnated with Cu2O. J. Photochem. Photobiol. A 2022, 427, 113775.
https://doi.org/10.1016/j.jphotochem.2022.113775.
65. Scherrer, P. Estimation of the size and internal structure of colloidal particles by means of röntgen. Nachr. Ges. Wiss. Göttingen
1918, 2, 96100.
66. Gao, L.; Luo, Y.; Kang, Y.; Gao, M.; Abdulhafidh, O. Experimental Study on Physical Mechanical Properties and
Microstructure of Diatomite Soil in Zhejiang Province, China. Appl. Sci. 2022, 12, 387.
67. Zuo, X.; Wang, L.; He, J.; Li, Z.; Yu, S. SEM-EDX studies of SiO2/PVDF membranes fouling in electrodialysis of polymer-
flooding produced wastewater: Diatomite, APAM and crude oil. Desalination 2014, 347, 4351.
https://doi.org/10.1016/j.desal.2014.05.020.
68. Sing, K.S.W. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area
and porosity (Recommendations 1984). Pure Appl. Chem. 1985, 57, 603619. https://doi.org/10.1351/pac198557040603.
69. Rouquerol, F.; Rouquerol, J.; Sing, K.S.W.; Llewellyn, P.; Maurin, G. Adsorption by Powders and Porous Solids, 2nd ed.; Academic
Press: Cambridge, MA, USA, 2014; https://doi.org/10.1016/C2010-0-66232-8.
70. Yurdakal, S.; Garlisi, C.; Özcan, L.; Bellardita, M.; Palmisano, G. Chapter 4(Photo)catalyst Characterization Techniques:
Adsorption Isotherms and BET, SEM, FTIR, UVVis, Photoluminescence, and Electrochemical Characterizations. In
Heterogeneous Photocatalysis; Marcì, G., Palmisano, L., Eds; Elsevier: Amsterdam, The Netherlands, 2019; pp. 87152.
https://doi.org/10.1016/B978-0-444-64015-4.00004-3.
71. Azimi Pirsaraei, S.R.; Asilian Mahabadi, H.; Jonidi Jafari, A. Airborne toluene degradation by using manganese oxide
supported on a modified natural diatomite. J. Porous Mater. 2016, 23, 10151024. https://doi.org/10.1007/s10934-016-0159-2.
72. Benkacem, T.; Hamdi, B.; Chamayou, A.; Balard, H.; Calvet, R. Physicochemical characterization of a diatomaceous upon an
acid treatment: A focus on surface properties by inverse gas chromatography. Powder Technol. 2016, 294, 498507.
https://doi.org/10.1016/j.powtec.2016.03.006.
73. Inchaurrondo, N.; Ramos, C.P.; Žerjav, G.; Font, J.; Pintar, A.; Haure, P. Modified diatomites for Fenton-like oxidation of
phenol. Microporous Mesoporous Mater. 2017, 239, 396408. https://doi.org/10.1016/j.micromeso.2016.10.026.
74. Mazidi, M.; Behbahani, R.M.; Fazeli, A. Screening of treated diatomaceous earth to apply as V2O5 catalyst support. Mater. Res.
Innovations 2017, 21, 269278. https://doi.org/10.1080/14328917.2016.1211478.
75. Xia, Y.; Li, F.; Jiang, Y.; Xia, M.; Xue, B.; Li, Y. Interface actions between TiO2 and porous diatomite on the structure and
photocatalytic activity of TiO2-diatomite. Appl. Surf. Sci. 2014, 303, 290296. https://doi.org/10.1016/j.apsusc.2014.02.169.
76. Sprynskyy, M.; Kowalkowski, T.; Tutu, H.; Cukrowska, E.M.; Buszewski, B. Ionic liquid modified diatomite as a new effective
adsorbent for uranium ions removal from aqueous solution. Colloids Surf. A Physicochem. Eng. Asp. 2015, 465, 159167.
https://doi.org/10.1016/j.colsurfa.2014.10.042.
77. Caliskan, N.; Kul, A.R.; Alkan, S.; Sogut, E.G.; Alacabey, İ. Adsorption of Zinc (II) on diatomite and manganese-oxide-
modified diatomite: A kinetic and equilibrium study. J. Hazard. Mater. 2011, 193, 2736.
https://doi.org/10.1016/j.jhazmat.2011.06.058.
78. Ilia, I.; Stamatakis, M.; Perraki, T. Mineralogy and technical properties of clayey diatomites from north and central Greece.
Open Geosci. 2009, 1, 393403. https://doi.org/10.2478/v10085-009-0034-3.
79. Kong, X.; Chong, X.; Squire, K.; Wang, A.X. Microfluidic diatomite analytical devices for illicit drug sensing with ppb-Level
sensitivity. Sens. Actuators B Chem. 2018, 259, 587595. https://doi.org/10.1016/j.snb.2017.12.038.
80. Rea, I.; De Stefano, L. Recent Advances on Diatom-Based Biosensors. Sensors 2019, 19, 5208.
81. Viji, S.; Anbazhagi, M.; Ponpandian, N.; Mangalaraj, D.; Jeyanthi, S.; Santhanam, P.; Devi, A.S.; Viswanathan, C. Diatom-
Based Label-Free Optical Biosensor for Biomolecules. Appl. Biochem. Biotechnol. 2014, 174, 11661173.
https://doi.org/10.1007/s12010-014-1040-x.
82. Rea, I.; Martucci, N.M.; De Stefano, L.; Ruggiero, I.; Terracciano, M.; Dardano, P.; Migliaccio, N.; Arcari, P.; Taté, R.; Rendina,
I.; et al. Diatomite biosilica nanocarriers for siRNA transport inside cancer cells. Biochim. Biophys. Acta Gen. Subj. 2014, 1840,
33933403. https://doi.org/10.1016/j.bbagen.2014.09.009.
83. Choudhary, S.; Sahu, K.; Bisht, A.; Singhal, R.; Mohapatra, S. Template-free and surfactant-free synthesis of CeO2 nanodiscs
with enhanced photocatalytic activity. Appl. Surf. Sci. 2020, 503, 144102. https://doi.org/10.1016/j.apsusc.2019.144102.
84. Kaviyarasu, K.; Fuku, X.; Mola, G.T.; Manikandan, E.; Kennedy, J.; Maaza, M. Photoluminescence of well-aligned ZnO doped
CeO2 nanoplatelets by a solvothermal route. Mater. Lett. 2016, 183, 351354. https://doi.org/10.1016/j.matlet.2016.07.143.
85. Wang, G.; Mu, Q.; Chen, T.; Wang, Y. Synthesis, characterization and photoluminescence of CeO2 nanoparticles by a facile
method at room temperature. J. Alloys Compd. 2010, 493, 202207. https://doi.org/10.1016/j.jallcom.2009.12.053.
Water 2022, 14, 3373 24 of 24
86. Parks, G.A. Aqueous Surface Chemistry of Oxides and Complex Oxide Minerals. In Equilibrium Concepts in Natural Water
Systems; Advances in Chemistry; American Chemical Society: Washington, DC, USA, 1967; Volume 67, pp. 121160.
87. Nosrati, A.; Larsson, M.; Lindén, J.B.; Zihao, Z.; Addai-Mensah, J.; Nydén, M. Polyethyleneimine functionalized mesoporous
diatomite particles for selective copper recovery from aqueous media. Int. J. Miner. Process. 2017, 166, 2936.
https://doi.org/10.1016/j.minpro.2017.07.001.
88. Sondi, I.; Bišćan, J.; Pravdić, V. Electrokinetics of Pure Clay Minerals Revisited. J. Colloid Interface Sci. 1996, 178, 514522.
https://doi.org/10.1006/jcis.1996.0146.
89. Ye, X.; Kang, S.; Wang, H.; Li, H.; Zhang, Y.; Wang, G.; Zhao, H. Modified natural diatomite and its enhanced immobilization
of lead, copper and cadmium in simulated contaminated soils. J. Hazard. Mater. 2015, 289, 210218.
https://doi.org/10.1016/j.jhazmat.2015.02.052.
90. Gao, B.; Jiang, P.; An, F.; Zhao, S.; Ge, Z. Studies on the surface modification of diatomite with polyethyleneimine and
trapping effect of the modified diatomite for phenol. Appl. Surf. Sci. 2005, 250, 273279.
https://doi.org/10.1016/j.apsusc.2005.02.119.
91. Sun, Z.; Yang, X.; Zhang, G.; Zheng, S.; Frost, R.L. A novel method for purification of low grade diatomite powders in
centrifugal fields. Int. J. Miner. Process. 2013, 125, 1826. https://doi.org/10.1016/j.minpro.2013.09.005.
92. Houas, A.; Lachheb, H.; Ksibi, M.; Elaloui, E.; Guillard, C.; Herrmann, J.M. Photocatalytic degradation pathway of methylene
blue in water. Appl. Catal. B 2001, 31, 145157.
93. Al-Qodah, Z.; Lafi, W.K.; Al-Anber, Z.; Al-Shannag, M.; Harahsheh, A. Adsorption of methylene blue by acid and heat treated
diatomaceous silica. Desalination 2007, 217, 212224. https://doi.org/10.1016/j.desal.2007.03.003.
94. Zhang, T.; Oyama, T.; Aoshima, A.; Hidaka, H.; Zhao, J.; Serpone, N. Photooxidative N-demethylation of methylene blue in
aqueous TiO2 dispersions under UV irradiation. J. Photochem. Photobiol. A 2001, 140, 163172. https://doi.org/10.1016/s1010-
6030(01)00398-7.
95. Zhang, T.; Oyama, T.k.; Horikoshi, S.; Hidaka, H.; Zhao, J.; Serpone, N. Photocatalyzed N-demethylation and degradation of
methylene blue in titania dispersions exposed to concentrated sunlight. Sol. Energy Mater. Sol. Cells 2002, 73, 287303.
https://doi.org/10.1016/s0927-0248(01)00215-x.
96. Lin, J.; Luo, Z.; Liu, J.; Li, P. Photocatalytic degradation of methylene blue in aqueous solution by using ZnO-SnO2
nanocomposites. Mater. Sci. Semicond. Process. 2018, 87, 2431. https://doi.org/10.1016/j.mssp.2018.07.003.
97. Hao, C.; Li, J.; Zhang, Z.; Ji, Y.; Zhan, H.; Xiao, F.; Wang, D.; Liu, B.; Su, F. Enhancement of photocatalytic properties of TiO2
nanoparticles doped with CeO2 and supported on SiO2 for phenol degradation. Appl. Surf. Sci. 2015, 331, 1726.
https://doi.org/10.1016/j.apsusc.2015.01.069.
98. Melcher, J.; Barth, N.; Schilde, C.; Kwade, A.; Bahnemann, D. Influence of TiO2 agglomerate and aggregate sizes on
photocatalytic activity. J. Mater. Sci. 2016, 52, 10471056. https://doi.org/10.1007/s10853-016-0400-z.
... The absorption changes of Rh B dye at maximum absorbance wavelength (max=553 nm) were monitored using a UV-vis spectrophotometer. Additional detailed information on the photocatalytic system was reported in our previous work (Turkten, 2022). ...
... Moreover, a small band at 587 cm −1 could be assigned to the stretching ( O-Ce-O) vibrational signal. The small bands located at 1042 cm −1 and 834 cm −1 belong to the stretching vibration of (NO3 -) indicating the existence of residual nitrate moiety(Ramadan & El-Masry, 2021;Turkten, 2022;Vivek & Babu, 2016;Xie et al., 2021). ...
Article
The discharge of untreated wastewater from unplanned industrial activities using dyes can cause serious environmental pollution and affect the aquatic environment. Semiconductor photocatalysis is a favorable technology widely used for degrading organic dyes in wastewater. This study dealt with the preparation of CeO2 nanoparticles via a simple precipitation technique. Information on the structural and morphological features of the developed CeO2 nanoparticles were determined using Fourier transform infrared with attenuated total reflectance (FTIR-ATR), Raman spectroscopy, X-ray diffraction (XRD), and scanning electron microscopy (SEM) spectroscopic methods. The presence of the characteristic bands of CeO2 in the FTIR spectrum provided evidence of successful CeO2 formation. The calculated crystallite particle size utilizing the Scherrer equation was 10 nm. SEM images revealed that the morphology of CeO2 consisted of almost spherical particles with slight agglomeration. Brunauer-Emmett-Teller (BET) technique was also used to find out the specific surface area of CeO2 nanoparticles (11 m2/g). The efficiency of CeO2 nanoparticles was also confirmed in terms of their photocatalytic activity against Rhodamine B (Rh B) under UV-A light. The results indicated that CeO2 nanoparticles could be a promising catalyst candidate for industrial wastewater treatment.
... Studies conducted using diatomite and hydroxyapatite as polyurethane reinforcements have shown that composites with diatomite have better thermal resistance than biocomposites with hydroxyapatite. Turkten [19] presented the developed CeO2diatomite composites with high photocatalytic performance, which be used to remove pollutants from water. ...
Article
Full-text available
Diatomite from a deposit in Jawornik Ruski (Poland) has been selected as the material for study. The paper aimeds to show the possibility of using diatomite from the Carpathian Foothills as a sorbent of petroleum substances.Diatomite in the delivery condition (DC) and diatomite after calcination were used for this study. The material was calcined at 600, 650, 750, 850 and 1000°C. The diatomaceous earth was then granulated. The morphology of diatomite was observed using SEM. Particle size distribution was determined by Laser Particle Analyzer, chemical composition was determined by XRF, and mineralogical composition by XRD. Specific surface area, pore volume and pore size were determined. Thermal analysis (TG, DTA) was carried out. Absorption capacity tests were performed and the effect of diatomite addition on water absorption of concrete samples was determined.Within the framework of the study, it was shown that diatomite from the Jawornik deposit could be successfully used as a sorbent for petroleum substances. The absorption capacity of calcined at 1000°C diatomaceous earth was 77%. The obtained result exceeds the effectiveness of previously used absorbents, for which the sorption level is 60-70%. This allows commercial use of diatomite from deposits in Poland. In addition, water absorption tests have shown that diatomaceous earth can successfully replace cement used in concrete productione. The most favourablee effect on the reduction of water absorption is the addition of diatomite in the amount of 10%.The properties of diatomaceous earth from the Jawornik Ruski deposit indicate its high potential for use in the synthesis of geopolymers, which is important not only from an economic but also from an ecological point of view.The novelty of this work is the demonstration of the possibility of using diatomite as a sorbent of petroleum substances with high efficiency, exceeding the previously used sorbents.
Article
Full-text available
Magnetite-diatomite nanocomposite was synthesized through co-precipitation methods as an effective Cr(VI) removal adsorbent. The properties of diatomite (DM), thermochemically modified diatomite (TMD), and magnetic-diatomite nanocomposite (MDM) were investigated using FTIR, X-ray diffraction analysis (XRD), BET, and complete silicate chemical analysis. The MDN shows 98.89% adsorption removal at optimized conditions using the response surface methodology (RSM) of Box–Behnken Design (BBD). The kinetic data for Cr(VI) sorption on MDN were well described by pseudo-second order, which indicates the Cr(VI) adsorption was mainly due to chemisorption. The isotherm data show that the Langmuir and Freundlich models better described Cr(VI) ion sorption data. The thermodynamic parameters ΔG°, ΔH°, and ΔS° were estimated, and the results indicate Cr(VI) sorption on MDN was a spontaneous (ΔG° < 0) and exothermic process (ΔH° < 0). The proper Fe3O4 loading into TMD improves the gram susceptibility (Xg) of MDMfor magnet separation. The regeneration of nanocomposite material revealed over 80% Cr(VI) removal efficiency after five consecutive adsorption–desorption cycles. The produced MDM was tested for the removal of Cr(VI) from real tannery wastewater. The obtained results suggest the possibility of using this nanocomposite as an effective, efficient adsorbent to remove Cr(VI) laden wastewater.
Article
Full-text available
In this study, the photocatalytic degradation of Methylene Blue was investigated using CdSe nanoparticles. CdSe nanoparticles were synthesized via a simple method and were characterized by FTIR, XRD, FESEM, BET, DRS and EDS techniques. The photocatalytic performance of the CdSe nanoparticles was optimized using Response Surface Methodology (RSM) under visible light. The independent variables involved initial pH, MB concentration, photocatalyst dosage, and irradiation time were evaluated and the optimum photodegradation efficiency of MB dye removal was achieved ˜ 92.80% at pH = 8, 20 mgL−1 of MB concentration, 0.02 g 50 mL−1 of CdSe dosage, and 20 min of irradiation time. Also, the photodegradation of MB by CdSe is obeyed pseudo-first-order kinetic model (k = 0.038 min−1). The thermodynamic results revealed that the photocatalytic degradation of MB is spontaneous and endothermic. Also, the evaluation of various scavengers confirmed that the MB photodegradation was mainly done by photogenerated holes and hydroxyl radicals.
Article
Full-text available
Although they are of significant importance for environmental applications, the industrialization of photocatalytic techniques still faces many difficulties, and the most urgent concern is cost control. Natural minerals possess abundant chemical inertia and cost-efficiency, which is suitable for hybridizing with various effective photocatalysts. The use of natural minerals in photocatalytic systems can not only significantly decrease the pure photocatalyst dosage but can also produce a favorable synergistic effect between photocatalyst and mineral substrate. This review article discusses the current progress regarding the use of various mineral classes in photocatalytic applications. Owing to their unique structures, large surface area, and negatively charged surface, silicate minerals could enhance the adsorption capacity, reduce particle aggregation, and promote photogenerated electron-hole pair separation for hybrid photocatalysts. Moreover, controlling the morphology and structure properties of these materials could have a great influence on their light-harvesting ability and photocatalytic activity. Composed of silica and alumina or magnesia, some silicate minerals possess unique orderly organized porous or layered structures, which are proper templates to modify the photocatalyst framework. The non-silicate minerals (referred to carbonate and carbon-based minerals, sulfate, and sulfide minerals and other special minerals) can function not only as catalyst supports but also as photocatalysts after special modification due to their unique chemical formula and impurities. The dye-sensitized minerals, as another natural mineral application in photocatalysis, are proved to be superior photocatalysts for hydrogen evolution and wastewater treatment. This work aims to provide a complete research overview of the mineral-supported photocatalysts and summarizes the common synergistic effects between different mineral substrates and photocatalysts as well as to inspire more possibilities for natural mineral application in photocatalysis.
Article
Full-text available
Due to the increased demand for clothes by the growing population, the dye-based sectors have seen fast growth in the recent decade. Among all the dyes, methylene blue dye is the most commonly used in textiles, resulting in dye effluent contamination. It is carcinogenic, which raises the stakes for the environment. The numerous sources of methylene blue dye and their effective treatment procedures are addressed in the current review. Even among nanoparticles, photocata-lytic materials, such as TiO2, ZnO, and Fe3O4, have shown greater potential for photocatalytic meth-ylene blue degradation. Such nano-sized metal oxides are the most ideal materials for the removal of water pollutants, as these materials are related to the qualities of flexibility, simplicity, efficiency, versatility, and high surface reactivity. The use of nanoparticles generated from waste materials to remediate methylene blue is highlighted in the present review.
Article
Full-text available
We herein aim to construct a novel sorbent (DT-S) possessing adsorption selectivity to mercury(II) in aqueous media. Silanized purified diatomite (DT-N) was first synthesized by grafting 3-aminopropyltrimethoxysilane (APS) onto purified diatomite (DT). Nanoporous DT-S was then constructed through successively grafting epichlorohydrin (ECH) and allyl thiourea (AT) onto DT-N. FTIR, elemental analysis (EA), Brunauer–Emmett–Teller (BET), XRD, SEM, and pH at the zero point of charge (pHzpc) results demonstrated that DT-S had ample –OH, –NH2, and C=S, specific surface area of 5.52 m²/g, small pore diameter (16.10 nm), porous structures, and pHzpc of 5.80, favorable for Hg(II) capture. Optimal adsorption parameters were determined through batch tests. Capture behavior was interpreted preferably by pseudo-second-order kinetic and Liu isothermal equations. DT-S’ capture features, e.g., monolayer, spontaneity, chemisorption, and endothermic reaction, were evidenced by the data obtained. DT-S had 56.30 mg/g of maximum adsorption capacity for Hg(II), exceeded some sorbents available. Competitive adsorption tests displayed remarkably selective ability to capture Hg(II) (> 80.01%). FTIR and XRD analyses validated a possible capture mechanism, i.e., chelation reactions took place between mercury ions and nitrogen, oxygen, or sulfur atom in solid–liquid interface. Altogether, DT-S with high removal efficiency, capture selectivity, and excellent reusability is expected to be new sorbent applied to Hg(II)-contaminated water purification. Graphical abstract
Article
Full-text available
To improve the photodegradation efficiency of organic dyes under visible light, a novel bio-inspired photo-catalyst was designed herein based on the theory of wood-bionics. Graded porous cerium oxide with wood structure (W-CeO 2) was prepared by using larch wood splinters as the template. W-CeO 2 maintains the structure of the wood. W-CeO 2 showed many advantages such as low consumption, high performance, and long-term stability for application as a photocatalyst. In this study, the morphological features, crystalline structure, specific surface area, chemical composition, carrier separation efficiency and optical properties of the as-developed photocatalyst were investigated comprehensively. The results show that W-CeO 2 completely replicated the multi-scale hierarchical pore structure of the wood template. Compared with cerium oxide prepared by the traditional hydrothermal method (0-CeO 2), the specific surface area of W-CeO 2 was doubled, and the average pore diameter was reduced by 73% to 5.38 nm. Moreover, W-CeO 2 showed the characteristics of anisotropy, circular electron diffraction pattern, reduced crystallinity, and higher charge separation efficiency. W-CeO 2 also possessed higher electronic activity, greater active oxygen content, and lower electron-hole recombination efficiency. Methylene blue (MB) dye was used as the target pollutant to test the photocatalytic performance of W-CeO 2. The MB degradation test results show that using wood as a template greatly improved the catalytic efficiency of the catalyst. The main free radicals involved in the reaction were determined by adding 1,4-benzo-quinone (BQ) and tert-butanol (TBA) for comparative experiments. The results of this work can provide new insights into the construction of bio-inspired materials for promoting the photocatalytic reaction of organic dyes and indicate that W-CeO 2 photocatalyst is a promising candidate for wastewater treatment application.
Article
Full-text available
In this paper, the solution combustion synthesis (SCS) accompanied with calcination at 500 and 600 °C (Ce1 and Ce-2) was used for the synthesis of cerium(IV) oxide (CeO2) nanoparticles as a facile and low-cost method using Ce(NO3)3∙6H2O as cerium precursor and salicylic acid as oxidizer and fuel. CeO2 nanoparticles were characterized by FT-IR and UV–Vis spectroscopies, VSM, XRD, EDS, BET and TEM. The FT-IR and XRD results confirmed the pure and single crystalline phase of CeO2 nanoparticles. The UV–Vis spectra of CeO2 nanoparticles predicted a narrow and uniform particle size distribution due to a strong band that appeared at ≈ 244 nm. The ferromagnetic nature of the as-prepared CeO2 nanoparticles was proved by VSM. Moreover, the photocatalytic degradation of methylene blue (MB) dye under UV irradiation was investigated and at optimum conditions, the MB removal percentage reaches 93 and 96% for Ce-1 and Ce-2, respectively.
Article
Full-text available
The green synthesis of metal oxides nanoparticles being as quick and ecofriendly synthetic method is a prevailing research trend alternative to toxic chemical synthetic routes. The objective of designing this research work is to reduce the negative impacts of synthetic methods and to flourish environmentally benign synthetic route for the metallic nanoparticles’ synthesis. The present study was carried out to bio-synthesize the MgO nanoparticles using a basic building block plant extract. The MgO nanoparticles were synthesized using aqueous Texas sage (Leucophyllum frutescens) plant extract solution as capping agent. The green route synthesized MgO nanoparticles were used for morphological studies by FESEM, TEM and EDX, Structural studies via XRD, XPS and FTIR and optical analysis by UV-DRS. The XRD confirms the cubic structure of the nanoparticles with size~34 nm. FT-IR analysis exposed the bio-chemical compounds responsible for the formation and ligation of the nanoparticles. TEM and FE-SEM revealed the rod-shaped nanoparticles formation. A characteristic metal and oxygen binding peak was observed at 296 nm that confirms the Mg-O formation. The photocatalytic effectiveness of synthesized nanoparticles, the degradation ability with was examined with Methylene blue (MB) under visible light irradiation. The 90% of degradation efficiency against MB dye was observed within 120 min of visible light irradiation. The antibacterial activity of the synthesized particles was tested against Staphylococcus aureus and Escherichia coli. It showed better bactericidal activity against Staphylococcus aureus as compared to Escherichia coli. The MgO nanoparticles are proved to be better and efficient applicant for water rescue and antibacterial materials. The overall results depicted that MgO nanoparticles obtained by green resource presented scope for multifaceted environmental and biological applications
Article
Full-text available
ZnO and TiO2 are semiconductor nanomaterials that are widely used in photocatalysis. However, the relatively high recombination rate and low quantum yield of photogenerated electron–hole pairs limit their practical applications. In this study, a series of TiO2/ZnO/diatomite composites with various compositions were successfully prepared via a two-step precipitation method. They exhibited stronger UV–visible absorption properties and substantially lower fluorescence intensities than those of ZnO and ZnO/diatomite, which was mainly due to the low recombination rate of the photogenerated electron–hole pairs in the composite system. The reaction intermediates of methylene blue were detected by liquid chromatography–mass spectrometry, and the degradation process was determined. The best composite catalyst was used for the degradation of gaseous methylbenzene and gaseous acetone. The gaseous acetone degradation product was determined to be acetaldehyde via gas chromatography–mass spectrometry. The results show that the composite catalyst exhibited a good photocatalytic degradation of both liquid pollutants and harmful volatile gases. When applied to the hydrogen and oxygen evolution reactions, the composite catalyst retained a good photoresponsivity and electrolytic efficiency.
Article
Full-text available
Over the last few years, various industries have released wastewater containing high concentrations of dyes straight into the ecological system, which has become a major environmental problem (i.e., soil, groundwater, surface water pollution, etc.). The rapid growth of textile industries has created an alarming situation in which further deterioration to the environment has been caused due to substances being left in treated wastewater, including dyes. The application of activated carbon has recently been demonstrated to be a highly efficient technology in terms of removing methylene blue (MB) from wastewater. Agricultural waste, as well as animal-based and wood products, are excellent sources of bio-waste for MB remediation since they are extremely efficient, have high sorption capacities, and are renewable sources. Despite the fact that commercial activated carbon is a favored adsorbent for dye elimination, its extensive application is restricted because of its comparatively high cost, which has prompted researchers to investigate alternative sources of adsorbents that are non-conventional and more economical. The goal of this review article was to critically evaluate the accessible information on the characteristics of bio-waste-derived adsorbents for MB’s removal, as well as related parameters influencing the performance of this process. The review also highlighted the processing methods developed in previous studies. Regeneration processes, economic challenges, and the valorization of post-sorption materials were also discussed. This review is beneficial in terms of understanding recent advances in the status of biowaste-derived adsorbents, highlighting the accelerating need for the development of low-cost adsorbents and functioning as a precursor for large-scale system optimization.
Article
One of the issues receiving global attention is the prevalence of dyes in wastewater and water resources. These chemicals' persistence, toxicity, and carcinogenic potential have a significant influence on aquatics and humans, making their removal prior to discharge into the environment necessary. Semiconducting metal oxide and modified metal oxide-mediated photocatalysis has emerged as a superior wastewater treatment process to tackle the major challenges faced by traditional technologies. The multifunctional ZnO nanomaterial due to its inexpensive nature, eco-friendly, structure-dependent properties, and complete mineralization of pollutants make them more efficient semiconducting photocatalysts than other materials. However, it has been seen that ZnO nanomaterials can show high performance in photo-induced degradation of cationic and anionic dyes and for the treatment of industrial effluent due to its high efficiency under ultraviolet light, with a 3.37 eV the bandgap. Hence this paper presents a review of recent achievements in the degradation of Methylene Blue, Rhodamine B as a cationic dye, and Methyl Orange, Eriochrome black T as an anionic dye, by studying the operating parameters reaction time, pH, initial dye concentration, and catalyst concentration.