1 s2.0 S0272884224033844 Main
1 s2.0 S0272884224033844 Main
Ceramics International
journal homepage: www.elsevier.com/locate/ceramint
A R T I C L E I N F O A B S T R A C T
Handling Editor: Dr P. Vincenzini The availability of fresh water is limited compared to its huge consumption. Conversely, certain industries like
dyes, textiles, paper, and tanning not only consume vast quantities of fresh water, but also generate significant
Keywords: amounts of wastewater that poses severe threats to living organisms. The present investigation focuses on the
Nanocomposites degradation of Rhodamine-B (Rh–B) as a model pollutant by using the nanocomposite based on waste scrap tire-
Activated carbon
derived activated carbon (AC) and titanium dioxide (TiO2). These nanostructures were fabricated using wet
TiO2
chemistry techniques. Analysis using scanning electron microscopy (SEM) revealed that the morphology of the
Photodegradation
Dye pollutants structures remained largely unchanged even with a concentration of 6 mg of AC. X-ray diffraction (XRD) results
confirmed the successful fabrication of rutile and anatase phases of TiO2@AC nanocomposite. The XPS results
confirmed the presence of AC and TiO2 in the composite. Furthermore, the optical band gap of the nano-
composites decreased up to 17.17 % and 34.57 % through direct and indirect methods respectively, owing to the
varying quantity of AC used. The narrowing of optical band gap and the increased surface oxygen vacancies were
caused by the successive addition of AC during the synthesis of nanocomposites. This resulted in the efficient
degradation of Rhodamine-B (Rh–B) using adsorption and photocatalytic oxidation. The findings demonstrate a
significant influence of AC on the TiO2 catalyst in the removal of Rh–B dye. The maximum efficiency of dye
removal, reaching up to 99.14 %, was achieved with the highest concentration of AC, thus demonstrating almost
complete elimination of Rh–B dye. The findings suggest that the use of waste scrap tires can be effectively applied
to design efficient photocatalysts for addressing the wastewater management issues.
1. Introduction colorants are heavily being used in industries like textile, paper, leather,
and tanning; and are benefiting the global economy. But these are very
Water shortage is a significant global concern that has become an harmful on the other hand. Therefore, getting rid of dye-contaminated
increasingly important issue in recent years. Besides household demands water is crucial owing to its certain causes, primarily arising from the
the major industry is the prime consumer of water. This consumption textile sector for instance. Most of the colorants used in the textile in-
has been increasing day by day with the rapid growth of industries and dustry are synthetic and are composed of aromatic rings in their mo-
population [1]. A wide range of pollutants, including fertilizers, pesti- lecular structure. These dyes are inert and non-biodegradable, making it
cides, textile dyes, and heavy metals, have been released into water challenging to clear them from wastewater without adequate treatment
streams as a result of industrial and agricultural activities [2]. Dye [3]. The incorporation of organic pollutants into water has a direct
* Corresponding author.
** Corresponding author.
E-mail addresses: [email protected] (S.N.U.S. Bukhari), [email protected] (A.A. Shah), [email protected] (W. Liu), iftikharc@
neduet.edu.pk (I.A. Channa), [email protected] (A.D. Chandio), [email protected] (I.A. Chandio), [email protected] (Z.H. Ibupoto).
https://doi.org/10.1016/j.ceramint.2024.07.440
Received 14 May 2024; Received in revised form 18 July 2024; Accepted 30 July 2024
Available online 2 August 2024
0272-8842/© 2024 Elsevier Ltd and Techna Group S.r.l. All rights are reserved, including those for text and data mining, AI training, and similar technologies.
S.N.U.S. Bukhari et al. Ceramics International 50 (2024) 41285–41298
impact on the food chain, leading to elevated levels of biochemical ox- in decomposing various organic pollutants than that of pure anatase or
ygen and affecting aquatic organisms [4]. Rh–B with the molecular rutile phases [15]. Besides, the phase formation of TiO2 nanoparticles
structure as shown in scheme 1b, is a commonly used fluorophore that and their short response in visible light and rapid recombination are
belongs to the xanthene dye family. It has various applications in in- significant drawbacks. To address these limitations, the addition of
dustries such as textile, cosmetic, laser, and pigment [5]. Unfortunately, carbon in the TiO2 nanoparticles can be a solution. Carbon can help
it is also one of the major pollutants that drain into water bodies, leading improve the photocatalytic activity of TiO2 by enhancing the separation
to serious health issues, including cancer [6]. Therefore, it is imperative and transfer of photogenerated electron–hole pairs. Carbon-based ma-
to focus on the degradation of Rh–B as a means of mitigating the terials have numerous applications across different fields due to their
problem of water pollution [7]. porous and large surface area characteristics [16]. Carbon materials are
Among various wastewater treatment techniques, the photocatalytic preferred over other materials due to their long-term stability, chemical
degradation process offers promising results [8] with the ability to compatibility, and mechanical and chemical properties. AC offers
degrade dyes economically without causing any further harm to the double-layered surface area and high porosity, making it suitable for gas
environment [9]. The use of TiO2 as a photocatalyst for the elimination and water separation and adsorption activity [17,18]. AC has both
of organic pollutants under ultraviolet (UV) or visible light irradiation mesoporous and microporous characteristics, allowing it to absorb
has become increasingly popular owing to its superior properties, molecules of varied sizes [19]. It can be synthesized from different waste
including high stability, strong redox ability, nontoxicity, good corro- materials such as waste plastics, bio-waste, and waste tires by pyrolysis
sion resistance, crystalline phase (anatase and rutile), pore size, aspect production process [20–22]. Moreover, the tire industry is a very
ratio, crystallinity, and low cost [10]. important part of the automotive industry and a huge contributor to the
TiO2, a widely used photocatalyst, exists in three distinct phases, global economy. But unfortunately, it is contributing the environmental
namely anatase, brookite, and rutile. Of these, rutile possesses a pollution as well. Scrap tires are a big environmental challenge with
tetragonal structure, while brookite has an orthorhombic structure [11]. disposal issues, non-biodegradable, accumulate hazardous gases
The rutile phase is considered to be the most thermodynamically stable [23–25]. The scrap tire is rich in carbon, and it can be used as a source of
among all the phases of titania [12,13]. The energy band gap of TiO2 is AC.
3.0 eV for rutile phase [14]. The mixture of the anatase and rutile phases However, the synthesis of AC from tire scrap and its combination
of TiO2 has been reported to exhibit higher photocatalytic performance with solvothermally grown TiO2 has not reported previously.
Scheme 1. (a) Detailed schematic diagram for synthesis of pure TiO2 nanoparticles and TiO2@AC nanocomposites using solvothermal method, (b) Molecular
structure of pollutant dye (Rh–B).
41286
S.N.U.S. Bukhari et al. Ceramics International 50 (2024) 41285–41298
Furthermore, we hereby show the effect of AC on narrowing the band stainless-steel autoclaves. The pre-annealed FTO substrates were then
gap of TiO2 and increasing the oxygen vacancies. Also an enhanced gently placed inside each Teflon bottle, with the conducting side facing
simultaneous adsorption and photocatalytic oxidation process towards upwards. The autoclaves were heated at 180 ◦ C for 40–60 min in a pre-
the removal of Rh–B was shown by the TiO2@AC nanocomposites. Such heated oven. In the last step, all FTO substrates were washed multiple
aspects of tire scrap for the preparation of AC and its combination with times in distilled water upon being removed from the autoclaves after
TiO2 has been rarely studied for degrading the Rh–B using both reaching room temperature. The product obtained from the FTO was
adsorption and photocatalytic oxidation process in existing literature. washed with double distilled water and filtered, then left to dry at room
In this study, the tire scrap was used to produce AC by using pyrolysis temperature for 18 h, as shown in Scheme 1a. Ultimately, pure TiO2
process. The different amounts of AC were further used during the sol- nanoparticles and TiO2@AC nanocomposites were obtained and stored
vothermal synthesis of TiO2 for developing the nanocomposites. The for further investigations and measurements.
structure, surface chemical composition and the optical properties of as Field Emission Scanning electron microscope (FESEM- Zeiss Supra
synthesized nanocomposites were studied. These nanocomposites were 55) was used to examine the morphology of the produced structures and
found highly active for the simultaneous degradation of Rh–B using EDS mapping was performed by XFlash Detector 630M (Bruker Ger-
adsorption and photocatalytic oxidation processes. The Rh–B removal many). The investigation of X-ray diffraction (XRD) patterns for the
percentage was found about 99.14 % from aqueous solution. synthesized samples was carried out using a German Bruker D8 Advance
X-ray diffractometer with Cu Kα radiation (λ = 0.15418 nm) as the
2. Material and methods source. The scattered radiation was detected within the angular range of
10–80◦ (2θ) with a scan rate of 0.02◦ . The AC altered TiO2 material’s
2.1. Materials surface was analyzed using Fourier Transform Infrared Spectroscopy to
identify the oxygenated functional group vibrational modes within the
All the necessary materials were procured from Sigma Aldrich in range of 400–4000 cm− 1. Before characterization, the samples were
Karachi, Sindh, Pakistan. These included titanium butoxide 97 % (Ti mixed with potassium bromide (KBr) at a ratio of 1:20. X-ray photo-
(C4H9O)4), n-butanone 99 % (C4H8O), hydrochloric acid 37 % (HCl), electron spectroscopy (XPS) measurements were obtained using a SHI-
fluorine doped titanium dioxide (FTO) glass substrate, acetone 99.5 % MADZU/Kartos Axis Supra system equipped with Al Kα radiation (hν =
(C3H6O), ethanol 99.5 % (C2H5OH), Rhodamine-B (Rh–B) 1486.6 eV). All samples were subjected to an electron take-off angle of
(C28H31ClN2O3), high purity nitrogen gas (N2) and Whatman filter 45◦ , and the C 1s peak level was utilized as an internal reference at
paper. All the reagents were of analytical grade and were used without 284.8 eV. It is noteworthy that this system boasts the ability to perform
any further alteration. parallel XPS imaging and scanning Auger microscopy (SAM) with
remarkable spatial resolutions of up to 1.0 and 0.1 μm, respectively.
2.2. Synthesis of activated carbon The exploration of optical properties holds immense significance in
the study of photo catalysis, as it provides a detailed account of the
Locally acquired discarded tires were ground and shifted into sizes of quantity of photons absorbed during the adsorption and photo catalytic
1–4 cm. After gentle washing, they were dried in an oven (DZF-6050-HT, process [26]. In order to obtain the optical properties of the synthesized
MTI Corporation) at a temperature of 110 ◦ C for a duration of 24 h samples in the range of 200–800 nm, the UV–Vis Spectroscopy was
before undergoing the pyrolysis process. The tire char, as the primary conducted using the state-of-the-art Shimadzu UV-2600
material, comprises 85 % of carbon by weight, which is sufficient for the spectrophotometer.
operation of a pyrolysis pilot plant. By subjecting the recycled tires to a The newly synthesized pure TiO2 and AC, TiO2@ 2 mg AC, TiO2@ 4
fixed average temperature of 450 ◦ C in the pyrolysis reactor, fuel gas mg AC, and TiO2@ 6 mg AC nanocomposites were employed to the
such as hydrocarbon fuel and tire char can be produced. To develop the photocatalytic degradation of the Rh–B dye, which is widely regarded as
activation process, nitrogen gas (99.7 % pure) is introduced into the a model wastewater pollutant. The photocatalytic activity was carried
furnace with a flow rate of 100 mL/30 s in order to eliminate other gases out under LED as a UV light source with a wavelength of 395 nm and a
from the furnace. Prior to the furnace reaching the desired activation power of 100 W (purchased from local market in Beijing), while varying
temperature range of 900–975 ◦ C in the presence of nitrogen gas, it is the time intervals from 0 to 150 min. A stock solution of 500 mL was
loaded with 30 g of tire char. When the required activation temperature created initially by mixing Rh–B dye at a concentration of 4.17 × 10− 5 M
is achieved, steam and nitrogen in a ratio of 85:15 are injected into the (20 mgL− 1). Five beakers were utilized in which 50 mL of the stock
activation furnace with a flow rate of 350 mL/30 s using a peristaltic solution was mixed with 5 mg (100 mg/L) of each synthesized nano-
pump. The activation process can run for a duration of 2–6 h in order to composite photo catalyst and bare TiO2 and AC. The resulting mixture
develop various porous structures. was mechanically dispersed in darkness for 90 min and kept on magnetic
stirrer, in order to achieve an equilibrium state of adsorption and
2.3. Synthesis of pure TiO2 nanoparticles and TiO2@AC nanocomposites desorption processes, during this the adsorption was continuously
measured after each 15 min intervals. The 5 mL of mixed solution was
The synthesis of pure TiO2 nanoparticles and TiO2@AC nano- taken from 50 mL beakers and centrifuged at 4000 rpm to separate the
composites was performed by a typical solvothermal method as catalyst from the solution. This process was performed with meticulous
described below: First, 1 mL of titanium butoxide was added into 10 mL care and precision to yield the best results. Prior to subjecting these
of ethanol in a beaker and sonicated for 20 min at room temperature, beakers to UV light irradiation for varying time periods in order to
resulting in the formation of a seed solution. The pre-cleaned FTO activate the photo catalysts and observe the degradation of the Rh–B,
substrates were then dipped into the seed solution and annealed in a pre- photo catalytic measurements were first taken through UV–Vis spec-
heated oven at 500 ◦ C for 20 min. Next, four beakers were taken for the trophotometer (PE Lamda365) to record the adsorption. The absorbance
preparation of a solution containing 1 mL of titanium butoxide, 6 mL of was adjusted to a wavelength of 553 nm, which is the optimal absorption
n-butanone, and 6 mL of hydrochloric acid at 37 % concentration, which point for the Rh–B dye and recorded. The adsorption of dye due to all
were then mixed by magnetic stirring for 20 min. The concentration of samples was measured in − 90 to 0 min the time to acquire adsorption/
titanium butoxide used for the growth solution was 0.23M. AC was desorption equilibrium. Then the photocatalytic activity for the same
added to three of the beakers in varying amounts of 2 mg, 4 mg, and 6 synthesized composites was estimated by measuring the decomposition
mg respectively, while the fourth beaker was kept as pure Titania. The rate of Rh–B in aqueous solution in the presence of UV light irradiation
resulting four beakers were named PT, TC-2 mg, TC-4 mg, and TC-6 mg. in time 0–150 min, and the results are shown in Fig. 8. The percentage of
Next, these four beakers were individually added in separate teflon-lined dye degradation was calculated using eq. (6) [27].
41287
S.N.U.S. Bukhari et al. Ceramics International 50 (2024) 41285–41298
Fig. 1. SEM image (a) pure TiO2, (b) 2 mg of AC assisted TiO2@AC (c) 4 mg of AC assisted TiO2@AC, (d) 6 mg of AC assisted TiO2@AC
41288
S.N.U.S. Bukhari et al. Ceramics International 50 (2024) 41285–41298
Fig. 2. EDS spectra and mapping (a) pure TiO2, (b) 2 mg of AC assisted TiO2@AC (c) 4 mg of AC assisted TiO2@AC, (d) 6 mg of AC assisted TiO2@AC
Fig. 3. (a) XRD patterns of pure TiO2 and, 2, 4, 6 mg of AC assisted TiO2@AC nanocomposites (b) Peak shift in two theta angles when AC was used.
41289
S.N.U.S. Bukhari et al. Ceramics International 50 (2024) 41285–41298
Table 1
Summary of some crystal parameters obtained from the pure TiO2 and TiO2@AC nanocomposite. (A and R stands for anatase and rutile phases respectively).
Dopant Crystallite Phase FWHM Peak Intensity D (nm) d (nm)
quantity A R A R A R A R A R
0 mg (101) (110) 0.248 0.908 346.48 318.14 32.47 8.9 3.481 3.231
2 mg (101) (110) 0.253 1.489 264.02 222.818 31.84 5.42 3.475 3.229
4 mg (101) (110) 0.241 1.405 362.919 289.58 33.42 5.75 3.477 3.229
6 mg (101) (110) 0.2 1.917 714.349 192.067 40.26 4.21 3.495 3.250
The average crystallite size of rutile phase is smaller (4.21 nm) in the 3437, 3416, and 3422 cm− 1, also the vibration peaks at 1751, 1781, and
6 mg of AC used for the synthesis of nanocomposites than the anatase 1791 cm− 1 were found to be attributed to the C–C bond and the
phase of TiO2 (40.26 nm) in the same nanocomposite as given in Table 1. stretching vibration of O–H bond, which could be caused by water
It can also be seen that the use of increasing amount of carbon molecules adsorbed on the surface of the samples [35,39,40]. Further-
contributed the increasing value of average crystallite size, however in more, the stretching modes of the –C=O group were also observed, as
the rutile phase the increasing content of AC resulted the decrease in the indicated by the peak at 1625 cm− 1, also this peak was linked to the C–O
average crystallite size. This is another aspect that the predominance of vibrations of the carboxyl groups in the acidic oxygenated surface [41].
rutile phase over anatase phase in terms of average crystallite size has These carboxyl groups play a crucial role in absorbing heavy metals, and
strongly supported the nanocomposites to act efficiently towards the their vibrations were indicative of the surface’s ability to absorb these
degradation of Rh–B. metals. The C–O bonds and aromatic ring stretching vibrations were
The FTIR spectra of the pure TiO2, TiO2@ 2 mg AC, TiO2@ 4 mg AC, found to be the sources of peaks at 1406, 1408, and 1418 cm− 1 [42].
and TiO2@ 6 mg AC nanocomposite are depicted in Fig. 4a-d. The results Absorption in the C–O stretching region at 1184 cm− 1 confirmed the
of pure TiO2 sample displays a range of vibrational peaks, with the most carbonyl band to an ester, while other bands were recorded between
prominent being 481, 584, 665, 880, 1086, 1189, 1418, 1504, 1634, 1100 cm− 1 and 1200 cm− 1 [43]. Furthermore, the bending mode of the
1781, 2338, 2464, 2992, 3392, 3724 and 3854 cm− 1. Meanwhile, the C–H or O–H was attributed to the peak at about 665 cm− 1[44]. This
TiO2@ 2 mg AC sample exhibits its own set of vibrational peaks, discovery may be attributed to the unique capability of TiO2 to absorb
including those at 526, 607, 832, 1092, 1184, 1418, 1514, 1625, 1791, molecules. Furthermore, the vibration peaks observed at 2947, 2952,
2320, 2464, 2947, 3437, 3722, and 3854 cm− 1. When it comes to the 2966, and 2992 cm− 1 from all samples consigned to C–H stretching vi-
TiO2@ 4 mg AC sample, the vibrational peaks present are at 602, 820, bration mode are indicative of the stretching modes of hydroxyl groups
1096, 1184, 1408, 1499, 1625, 1781, 2333, 2459, 2952, 3416, 3728 and (O–H) [45]. These peaks were observed in all samples, indicating that
3849 cm− 1. Finally, the TiO2@ 6 mg AC sample showcases its own the hydroxyl group is a common component in these samples. In all
unique range of vibrational peaks, with the most prominent being at samples, stretching modes of the hydroxyl group (O–H) can be observed
525, 826, 1096, 1194, 1406, 1613, 1751, 2243, 2382, 2966, 3422, 3773, from 3700 to 3800 cm− 1 [35]. These observations imply that the hy-
and 3880 cm− 1 this all information is shown in Fig. 4a-d, respectively. droxyl group is present in all the samples, which may have been
The wide band in the 400-880 cm− 1 range, which can be attributed to absorbed on the surface of TiO2. These findings suggest that TiO2 may
the stretching vibration of the Ti–O and Ti–O–Ti bond, is responsible for have unique adsorption properties.
the characteristic peaks of TiO2 [35,36]. It indicates the presence of a The FTIR spectra have provided evidence of the presence of oxygen-
mixture of polymorphic phases of anatase and rutile, as observed in the containing functional groups in the activated carbon doped TiO2
range of 480–880 cm− 1, which is consistent with previous studies. This nanocomposites. It has been observed that the prepared material pos-
phenomenon is related to the formation of TiO2 and the dopant’s sesses additional active sites, which augment its adsorption abilities. The
incorporation into the TiO2 lattice, as evidenced by the observed shift in formation of AC-loaded TiO2 nanocomposites has been established
the composite spectra [37,38]. The peaks observed in the infrared through the variation in peaks in all samples. Oxygen–metal–oxygen
spectra of the samples, namely at 1634 and 1613 cm− 1, as well as 3392, bonding has been confirmed through the analysis of the aforementioned
peaks [46,47]. Furthermore, the AC loading has significantly modified
the adsorption capability of TiO2 nanoparticles. It has been observed
that both surface hydroxyl groups and water molecules can react with
photo-generated holes to produce active hydroxyl radicals. This, in turn,
contributes to the enhancement of photo-activity [48]. The observations
confirm that photocatalytic reactions have taken place on the surface of
newly synthesized nanocomposite samples [49]. Furthermore, the
addition of AC has slightly changed the vibrational bands, thus these
were possibly caused by the interaction of AC with TiO2 as shown in
Fig. 4a–d.
X-ray photo-electron spectroscopy, a technique used to study the
composition of synthesized nanocomposites at a surface elemental level,
was utilized in the research. This technique offers a precise numerical
description of the chemical composition of a material surface within a
range of 10 nm and can detect elemental information with high sensi-
tivity ranging from 0.1 to 1 at % for all elements excluding H and He,
thus making it a valuable method for analyzing surface properties [50,
51]. The deconvolution of the XPS spectra was carried out using
Gaussian functions. S1(a, b) presents a comprehensive XPS survey
spectrum of pure TiO2 and TiO2@ 2 mg AC. The results demonstrated
that the synthesized samples were made up of C, O, and Ti solely. The
binding energies for C 1s, O 1s, and Ti 2p3/2 were determined to be
Fig. 4. FTIR spectra (a) pure TiO2, (b) 2 mg of AC assisted TiO2@AC (c) 4 mg of
AC assisted TiO2@AC, (d) 6 mg of AC assisted TiO2@AC 284.79 eV, 529.86 eV, and 458.43 eV, respectively, for pure TiO2. The
41290
S.N.U.S. Bukhari et al. Ceramics International 50 (2024) 41285–41298
binding energies for C 1s, O 1s, and Ti 2p3/2 were 284.79 eV, 529.57 eV, indicated that the oxygen species in the TiO2 lattice were substituted by
and 458.43 eV, respectively, in the case of TiO2@ 2 mg AC. Additionally, carbon species in the samples [54,55]. Such substitution can further
S1(c, d) represents a full XPS survey spectrum of TiO2@ 4 mg AC and enhance the visible light response of TiO2.
TiO2@ 6 mg AC. The binding energies for C 1s, O 1s, and Ti 2p3/2 were The introduction of carbon doping is known to have an impact on the
determined to be 284.75 eV, 529.7 eV, and 458.45 eV, respectively, for binding energy of O 1s, as evidenced by the deconvolution of the O 1s
TiO2@ 4 mg AC. The binding energies for C 1s, O 1s, and Ti 2p3/2 were spectrum for TiO2 at varying levels of AC concentration. Fig. 5a-d illus-
284.76 eV, 529.64 eV, and 458.41 eV, respectively, in the case of TiO2@ trates the deconvolution results for pure TiO2, TiO2@ 2 mg AC, TiO2@ 4
6 mg AC. These findings indicate that the synthesized nanocomposites mg AC, and TiO2@ 6 mg AC. The fitted data for the O 1s spectrum of
were made up of a mixture of carbon, oxygen, and titanium elements. pure TiO2 showed three distinct peaks at 529.57 eV, 531.17 eV, and
The binding energies for each element were slightly different because of 532.52 eV. Meanwhile, the O 1s spectrum of TiO2@ 2 mg AC exhibited
the variable concentration of AC used in the mixture. The XPS survey three peaks located at 529.86 eV, 531.62 eV, and 533.81 eV. For TiO2@
spectrum of the synthesized nanocomposites enabled us to gain a better 4 mg AC, the fitted data for the O 1s spectrum yielded three peaks at
understanding of the bonding and chemical state of the elements in the 529.7 eV, 531.67 eV, and 533.67 eV.
nanocomposites. Similarly, the O 1s spectrum of TiO2@ 6 mg AC also showed three
In the pursuit of examining the carbon states in Pure TiO2 and TiO2@ peaks located at 529.64 eV, 531.76 eV, and 533.49 eV. The binding
2 mg AC, the C 1s core level was measured, and the analysis was pre- energies at 531.17 eV and 532.52 eV for TiO2@ 0 mg AC, 531.67 eV and
sented in S2(a, b). The C 1s spectrum for Pure TiO2 was fitted into three 533.67 eV for TiO2@ 4 mg AC, and 531.76 eV and 533.49 eV for TiO2@
peaks positioned at 284.79 eV, 286.17 eV, and 288.83 eV. On the other 6 mg AC are associated with the Ti–O and CO (COO and C–OH) species,
hand, the C 1s spectrum for TiO2@ 2 mg AC nanocomposites was fitted respectively [56,57]. Furthermore, the broadening observed around
into four peaks located at 283.45 eV, 284.79 eV, 286.32 eV, and 288.81 532.52 eV, 533.67 eV, and 533.49 eV may be attributed to the presence
eV. The C 1s spectra for TiO2@ 4 mg AC and TiO2@ 6 mg AC were of surface hydroxyl groups, which are typically present on the surface of
represented in S2 (c, d). The C 1s spectrum for TiO2@ 4 mg AC was fitted nano crystalline TiO2 [58]. Moreover, XPS study has shown that the 6
into four peaks located at 283.42 eV, 284.75 eV, 286.2 eV, and 288.75 mg of AC assisted TiO2 nanocomposite exhibited higher amount of ox-
eV. Similarly, the C 1s spectrum for TiO2@ 6 mg AC nanocomposites was ygen vacancies which could play significant contributing role in
also fitted into four peaks located at 283.42 eV, 284.76 eV, 286.29 eV, enhancing the photodegradation of Rh–B.
and 288.68 eV. The peaks at 284.79 eV, 284.75 eV, and 284.76 eV for The O 1s binding energies also showed a slight positive shift from
nanocomposites with AC doping of 2 mg, 4 mg, and 6 mg, respectively, 531.17 eV for Pure TiO2 to 531.62 eV for TiO2@ 2 mg AC, 531.67 eV for
can be attributed to the C–C bond from the residual elemental carbon TiO2@ 4 mg AC, and 531.76 eV for TiO2@ 6 mg AC. This shift suggests
from the precursor or adventitious carbon present in the samples. These the formation of oxygen vacancies in the TiO2 lattice [50].
carbon species can act as photosensitizers that can enhance the visible Fig. 6a- d, demonstrate the de-convolution of the Ti 2p spectrum for
light response of TiO2 [52,53]. Furthermore, the peaks at 286.17 eV, various concentrations of activated carbon (AC) doped onto titanium
286.2 eV, and 286.29 eV for nanocomposites with AC doping of 2 mg, 4 dioxide (TiO2). These concentrations include pure TiO2, TiO2@ 2 mg
mg, and 6 mg, respectively, could be ascribed to the C–O–Ti bond of the AC, TiO2@ 4 mg AC, and TiO2@ 6 mg AC. The Ti 2p3/2 and Ti 2p1/2
synthesized nanocomposites. The presence of peaks at 283.46 eV, binding energy levels for pure TiO2 were seen at 458.46 eV and 464.33
283.42 eV, and 283.42 eV in 2 mg, 4 mg, and 6 mg, respectively, eV, respectively. The de-convolutions of Ti 2p for TiO2@ 2 mg AC
Fig. 5. XPS O1s spectra (a) pure TiO2, (b) 2 mg of AC assisted TiO2@AC (c) 4 mg of AC assisted TiO2@AC, (d) 6 mg of AC assisted TiO2@AC
41291
S.N.U.S. Bukhari et al. Ceramics International 50 (2024) 41285–41298
Fig. 6. XPS Ti 2p spectra (a) pure TiO2, (b) 2 mg of AC assisted TiO2@AC (c) 4 mg of AC assisted TiO2@AC, (d) 6 mg of AC assisted TiO2@AC
occurred at binding energy levels around 458.45 eV (Ti 2p3/2) and other stage is the 150 min UV light irradiation, which is mainly caused
464.27 eV (Ti 2p1/2), while for TiO2@ 4 mg AC, they occurred around by the photocatalytic degradation as shown in Table 2. In this work, the
458.43 eV (Ti 2p3/2) and 464.22 eV (Ti 2p1/2). Similarly, for TiO2@ 6 mg concentration of Rh–B maintained in the aqueous solution at the end of
AC, the de-convolution of Ti 2p occurred at binding energies around adsorption stage in the dark is taken as the starting concentration of
458.41 eV (Ti 2p3/2) and 464.17 eV (Ti 2p1/2), which can be associated Rh–B in the photocatalytic degradation stage, and the degradation
with the spin-orbit coupling of Ti 2p3/2 and Ti 2p1/2. The spin-orbit percentages of AC, pure TiO2 and TiO2@AC composites were calculated
splitting between Ti 2p3/2 and Ti 2p1/2 was calculated to be 5.87 eV, according to this reference. It is observed that dye removal is continue
5.82 eV, 5.79 eV, and 5.76 eV for pure TiO2, TiO2@ 2 mg AC, TiO2@ 4 due to AC in the presence of UV light up to 66.23 %, this shows that AC
mg AC, and TiO2@ 6 mg AC, respectively. These values indicate a has ability to adsorb dye up to time 150 min. Whereas the increase in dye
valence state of +4 for Ti, which corresponds to the tetravalent state of removal percentage in time 0–150 min was observed in the presence of
Ti [46]. Incorporation of carbon resulted in a shift of the XPS spectrum TiO2 due its photocatalytic characteristics, a 58.07 % of dye removal is
towards lower binding energies of AC doped nanocomposites (458.45 observed. While the increase in dye removal in the presence of TiO2@AC
eV, 458.43 eV, and 458.41 eV) in lieu of 458.46 eV (pure TiO2). The composites was observed due to both effects adsorption due to AC and
decrease in binding energy observed can be ascribed to the partial photocatalytic absorption due to TiO2. As shown in Fig. 7 (f), the initial
transfer of charges from the functional groups present on the carbon removal percentage of TiO2@AC composites was approximately similar
surface to the Ti4+ centers in TiO2. In addition, the decline in the binding to AC, but, after 90 min in the dark, the TiO2@AC composites are still
energies of AC doped nanocomposites in comparison to pure TiO2 can be able to remove the Rh–B by the photocatalytic degradation effect on the
credited to the intense interplay between Ti4+ and carbon elements, UV light irradiation; as a result, the percentages of Rh–B removal
which display lattice distortions of TiO2 [59]. continue to increase. The final percentages of Rh–B removed by TiO2@
2 mg AC, TiO2@ 4 mg AC, and TiO2@ 6 mg AC reach 83.39, 94.09 and
99.14 % respectively. Based on the results that the percentages of Rh–B
3.2. Photocatalytic analysis of as synthesized nanocomposites
removal due to AC, pure TiO2, TiO2@ 2 mg AC, TiO2@ 4 mg AC, and
TiO2@ 6 mg AC nanocomposites are 60.3 %, 0.42 %, 28.33 %, 39.98 %,
To explore the impact of the concentration of dopant (AC), a series of
and 55.14 % respectively after 90 min in the dark by adsorption, the
samples were examined by UV–vis spectrometer. For the pure AC
photocatalytic degradation percentage of Rh–B is 14.93 %, 57.89.5 %,
(Fig. 7a), the removal of Rh–B increases gradually with the increasing
76.82 %, 90.15 % and 98.08 % for AC, pure TiO2, TiO2@ 2 mg AC,
adsorption time, and about 63.3 % of the initial Rh–B is removed until
TiO2@ 4 mg AC, and TiO2@ 6 mg AC nanocomposites respectively.
reaching the adsorption equilibrium at the end of 90 min in the dark.
Clearly, the photocatalytic degradation percentage of Rh–B for synthe-
The Rh–B removal rate of pure TiO2 is negligible when the solution is
sized pure TiO2, TiO2@ 2 mg AC, TiO2@ 4 mg AC, and TiO2@ 6 mg AC is
kept in the dark for 90 min (Fig. 7 b) but increases noticeably with
increasing indicating that the photocatalytic activity of synthesized
prolonging the time of UV light irradiation, and the final removal rate of
TiO2@ 2 mg AC, TiO2@ 4 mg AC, and TiO2@ 6 mg AC is higher than that
Rh–B reaches only 58.07 %, which is mainly attributed to the photo-
in pure TiO2. The fundamental mechanism for this can be clarified as
catalytic degradation process. However, for the TiO2@AC composites
follows: the TiO2@AC composites exhibit the dual functions such as
(Fig. 7c–e), the Rh–B removal process in aqueous solution can be divided
absorption and photocatalytic degradation to remove the Rh–B in the
into two stages: the removal process of 90 min in the dark, which is
aqueous solution [60,61]; the proposed schematic illustration is shown
mainly attributed to the adsorption effect of TiO2@AC composites. The
41292
S.N.U.S. Bukhari et al. Ceramics International 50 (2024) 41285–41298
Fig. 7. UV–Vis spectrum for photodegradation of Rh–B dye by using various newly synthesized photocatalysts (a) bare AC, (b) pure TiO2, (c) 2 mg of AC assisted
TiO2@AC (d) 4 mg of AC assisted TiO2@AC, (e) 6 mg of AC assisted TiO2@AC and (f) their corresponding Rh–B removal % at − 90 to 150 min.
Table 2
Summarized results of UV–vis spectrometer for removal of Rh–B dye.
3
Sample Band gap (eV) Rate constant (10− min− 1) Adsorption efficiency (%) Photocatalytic efficiency (%) Overall dye removal % (− 90 to 150 min)
(-90 to 0 min) (0–150 min)
in Scheme 2, which is also similar to the case on the photocatalyst of with Rh–B molecules and there will more photocatalytic activity of
TiO2@AC nanocomposites. Under dark conditions, the removal process TiO2@AC nanocomposites particles. The study’s results indicated that
depends strongly on the adsorption performance of substrate AC. the newly synthesized nanocomposites had the greatest photocatalytic
However, under UV light irradiation, the Rh–B removal process requires activity in Rh–B dye degradation due to their dual phase (anatase and
the pollutant to be adsorbed on the TiO2@AC composites surface prior to rutile) with addition of a significant amount of AC as dopant. The TiO2@
immediate photocatalytic degradation [62,63]. Therefore, the efficiency 6 mg AC nanocomposite achieved the highest dye degradation at 99.14
of photocatalytic reaction can be enhanced by the adsorption perfor- %, owing to its significant presence of simultaneous adsorption and
mance of substrate AC as there is more Rh–B molecules attracted to- photocatalytic oxidation process as depicted in Table 1. Additionally,
wards TiO2@AC nanocomposites so there is more interaction of TiO2 studies have shown that a combination of anatase and rutile phases is
41293
S.N.U.S. Bukhari et al. Ceramics International 50 (2024) 41285–41298
Scheme 2. Schematic diagram of photocatalytic mechanism of bare AC, pure TiO2, AC assisted TiO2@AC photocatalysts (a) Adsorption (b) Photocatalysis.
more active than pure anatase/rutile [64]. The superior effectiveness is TiO2@ 2 mg AC, TiO2@ 4 mg AC, and TiO2@ 6 mg AC and AC were
attributed to the multiphase characteristic of the particles, which en- found to have rate constants of 0.048, 3.87, 5.73, 9.4 and 10.86 min⁻1,
hances the separation of electron-hole pairs and reduces the recombi- respectively for adsorption in dark and the photocatalysts pure TiO2,
nation rate of electron-hole pairs [65]. TiO2@ 2 mg AC, TiO2@ 4 mg AC, and TiO2@ 6 mg AC and AC were
The kinetics of Rh–B dye degradation were examined using a variety found to have rate constants of 6.07, 10.76, 15.54, 26.39 and 1.03 min⁻1,
of recently synthesized photocatalysts. The calibration curves were respectively for absorption in UV light irradiations (Table 2. And Fig. 8(a
generated to show the photo degradation efficiency (%), as well as the and b)). A higher rate constant implies a higher catalytic activity effi-
time interval of − 90 to 0 in dark and 0–150 min in UV light irradiations ciency [67]. These results provide valuable insights into the perfor-
as shown in Fig. 7f. Additionally, a plot of Ct/Co was created to represent mance of the photocatalysts and their potential applications in various
the result of adsorption and photo degradation of Rh–B dye at − 90 to 0 fields.
min and 0–150 min respectively. Finally, a pseudo-first order (ln Ct/Co The injection of electrons in the conduction band of pure TiO2 and
versus dark/irradiation time) plot was analyzed to further understand AC doped TiO2 was favored by the UV irradiation of light over the
the adsorption/photo degradation of Rh–B dye at given time intervals, synthesized catalyst, through surface phonon resonance mechanism,
which shows the photocatalytic activity increases as the irradiation time which aided in the production of free radicals [66]. The exceptional
increases, as demonstrated in Fig. 8(a and b). This increase in degra- photocatalytic performance of TiO2 based nanocomposites is primarily
dation efficiency is due to the contribution of adsorption incorporation attributed to their small size, extensive specific surface area, phase pu-
with photodegradation. It is important to note that the elimination of rity, and pristine surfaces. The large surface area and uncontaminated
dye color may occur due to the breakdown of chromophores that are surface of TiO2 nanoparticles are thought to aid in the adsorption of
available in the dye, which ultimately give the dye its specific color. By Rh–B, resulting in a higher rate of photo-excited electron transfer from
utilizing the Langmuir Hinshelwood (LH) model in eq. (4), the elimi- Rh–B to photocatalyst and enhanced dye degradation efficiency. The
nation rate constant of Rh–B was determined [66], results are mentioned free radicals exhibit a substantial capacity for oxidation, enabling them
in Table 2. to eliminate detrimental organic pollutants such as carboxylic acids
( ) [68].
Ct
ln = Kapp (4) Scheme 2 depicts the schematic mechanism of the photocatalytic
Co
reaction. The corresponding equations for the photodegradation mech-
anism of the organic pollutants on the Photocatalyst under UV light are
Where the initial concentration of the Rh–B dye is represented by Co,
presented in the text [69].
while the concentration after regular intervals of time is denoted as Ct.
The rate constant, k, can be determined from the slope of the irradiation TiO2/TiO2@AC (Photocatalyst) + hν → e− + h+ …. …. …. …. …. ….
time graph of ln Co/Ct. The linearity of all the curves indicates the …... (5)
presence of pseudo first order kinetics. The photocatalysts pure TiO2,
Fig. 8. Pseudo-first order plot for photodegradation of Rh–B dye at − 90 to 150 min by using various newly synthesized photocatalysts (a) bare AC, pure TiO2, 2 mg
of AC assisted TiO2@AC, 4 mg of AC assisted TiO2@AC, 6 mg of AC assisted TiO2@AC, (b) Plot of Ct/Co verses irradiation of photodegradation of Rh–B dye at − 90 to
150 min for the illustration of degradation kinetics.
41294
S.N.U.S. Bukhari et al. Ceramics International 50 (2024) 41285–41298
H2O/OH− + h+ → •OH + H+ …. …. …. …. …. …. …. …. …. …. …. ….
…. …. …. (6) portionality constant (band tailing parameter), hν denotes the incident
photon energy, and Eg is the band gap energy. The value of n, which
O2 + e →−
O2
• −
…. …. …. …. …. …. …. …. …. …. …. …. …. …. …. ….
varies depending on the type of optical transition of semiconductor, is
…. …. …. …. (7)
equal to 1 for direct transition and four for indirect [73].
O•-
2 + H → OOH …. …. …. …. …. …. …. …. …. …. …. …. …. …. …. ….
+ • The absorption curve for pure and altered TiO2 structures is illus-
…. …... (8) trated in Fig. 9a. By plotting the graph (αhν)2 verses hν and (αhν)1/2
verses hν for direct and indirect band gap, the reduction in band gaps up
O•-
2 + H + OOH → H2O2 + O2 …. …. …. …. …. …. …. …. …. …. …. ….
+ •
to 17.17 % and 34.57 % was observed for direct and indirect band gap
…. (9) respectively, results can be seen in Table 2 and Fig. 9(b and c). The
process of measuring band gap energy is crucial for the development of
H2O2 + O•-
2 → OH + OH + O2 …. …. …. …. …. …. …. …. …. …. …. ….
• −
materials with specific properties. As materials with smaller band gap
…. (10)
energy have more valence electrons available for conductance, they
•
OH + O•-
2 + Rh–B (C28H31ClN2O3) → CO2+H2O … … … …. …... (11) have enhanced electrical properties. The UV. Vis spectrometer gives the
absorbance curves with respect to wavelength and indicate the absor-
The said equations provide a comprehensive understanding of the bance edges at 340, 344, 353, and 360 nm for newly synthesized pure
process involved in the removal of Rh–B dye. The photodegradation TiO2, TiO2@ 2 mg AC, TiO2@ 4 mg AC, and TiO2@ 6 mg AC nano-
mechanism of the organic pollutants on the Photocatalyst under UV light composites respectively. These edges are indicative of the energy
is a crucial process that has been thoroughly studied, and its efficiency required to move an electron from the valence band to the conduction
has been demonstrated. The process described above results in the cre- band, and thus the band gap energy. The absorbance spectra also enable
ation of radicals that possess potent oxidation capabilities. When these the determination of the optical properties of the material, such as
radicals come into contact with organic pollutants, they break them refractive index and extinction coefficient. The information obtained
down into simpler and more harmless substances like carbon dioxide, from these measurements is useful for designing and optimizing devices
water, and other mineralized products. When AC doped TiO2 nano- that rely on the optical and electrical properties of the material. The
composites are exposed to UV light, the electrons present in the valence absorbance spectra of these nanocomposites indicate that the addition of
band become excited and move to the conduction band, leaving a hole AC affects the band gap energy of the material. The comparison of rate
behind in the valence band. These newly generated charge carriers then constant for adsorption and absorption of all synthesized samples is
engage in redox reactions and go on to oxidize oxygen molecules into shown in Fig. 9 d, which reveals that rate constant for adsorption is
superoxide ions. Concurrently, the water molecules present are reduced higher in samples containing higher quantity of AC due to adsorbing
to hydroxyl radicals. These radicals are incredibly reactive and when quality of AC [74], and rate constant for absorption is higher for samples
they encounter organic pollutants, they break them down into simpler, containing higher quantity of AC [75], as decrease in band gap energy is
less harmful substances [26,70]. For a synthesized photocatalyst to be observed in higher concentrated samples like in TiO2@ 6 mg AC sample.
effective, various interfacial electron processes involving e− and h+ must The use of nanocomposites, such as TiO2@AC, is becoming increasingly
be able to compete successfully with major deactivation processes popular due to their unique properties and potential applications
involving e− and h+ recombination. These deactivation processes can because of their admirable adsorption and absorption qualities [76].
occur either in the bulk or at the surface (as shown in Scheme 2). In our This study suggests that the addition of AC alters the physical and
case the presence of AC adsorbs the dye molecules due to their enter chemical properties of the material, and further research in this area is
molecular attractive forces, which causes to attract more dye molecules warranted, which could be useful for developing new materials with
towards the photocatalysts and being near to the surface of photo- enhanced performance and achieving the desired properties.
catalysts dye molecules have more interactions and in the result more Finally, a comparison is made between the results of the photo cat-
dye degradation occurs. If the inter facial electron processes can effec- alytic experiments using different catalysts available in the literature
tively compete with the major deactivation processes, the synthesized [77–91]. These experiments focus on the degradation of Rh–B dye, as
photocatalyst will be able to effectively break down organic pollutants listed in Table 3. It is evident from the results that this study shows great
and render them harmless. promise in terms of activity compared to other studies, particularly in
The ideal photocatalyst should possess several characteristics, such the use of UV light irradiation. Furthermore, the straightforward syn-
as being chemically and biologically inert, photo catalytically stable, thesis process and the enhancement of the photocatalytic process sug-
easy to produce and use, and efficient in catalyzing reactions. Addi- gest that the nanocomposites produced may have potential applications
tionally, it should be activated by UV/sunlight, inexpensive, and safe for in areas such as dye sensitized solar cells, adsorption, and gas sensors.
the environment and humans. Therefore, based on the results presented in Tables 3 and it can be
Researchers suggest that anatase’s increased photo reactivity is due inferred that the TiO2@AC photocatalysts prepared in this study exhibit
to its slightly higher Fermi level, lower oxygen adsorption capacity, and exceptional catalytic performance in the removal of Rh–B from indus-
higher degree of hydroxylation, i.e., the number of hydroxy groups on trial wastewater.
the surface [71,72].
The measured band gap of the samples was prepared with utmost
4. Summary
care and precision by dispersing synthesized, Pure TiO2 and AC assisted
TiO2@AC nanocomposite photocatalysts in ethanol. The absorption
Activated carbon was successively derived from waste material of
peak was at 283 nm due to TiO2 nanoparticles and there was no sig-
tire scrap and conjugated with TiO2 fabricated by simple solvothermal
nificant difference in absorption spectra of synthesized nanocomposites.
technique. The produced nanocomposite was successively applied to
The band gap of newly synthesized nanocomposites was determined
photocatalytic elimination of Rh–B dye chosen as model pollutant. The
with the help of the celebrated Tauc eq. (12) [72]. It is a scientific
SEM and EDS results confirmed the formation of nanorods and presence
method that has been extensively used to calculate the energy band gap
of respective elements (Ti, O and C) respectively and the elemental
of various materials, including semiconductors and insulators, with
presence was confirmed by XPS. FTIR results demonstrated the presence
remarkable accuracy.
of active special sites for the enhancement of degradation efficiency. The
(αhν) 2/n = A (hν − Eg) (12) newly synthesized nanocomposites have a remarkable performance in
photocatalysis. Moreover, the presence of AC in TiO2 resulted in an in-
In this equation, α stands for the absorption coefficient, A is the pro- crease photocatalytic elimination of Rh–B with a maximum degradation
of 99.14 % for 6 mg of AC addition into TiO2. Furthermore, the effect of
41295
S.N.U.S. Bukhari et al. Ceramics International 50 (2024) 41285–41298
Fig. 9. UV–vis absorbance spectra (a)) pure TiO2, 2 mg of AC assisted TiO2@AC, 4 mg of AC assisted TiO2@AC, 6 mg of AC assisted TiO2@AC and (b, c) their
corresponding direct and indirect band gap analysis using Tauc’s plot (d) The rate constants for adsorption and absorption processes using newly prepared pho-
tocatalysts such as bare AC, pure TiO2, 2 mg of AC assisted TiO2@AC, 4 mg of AC assisted TiO2@AC, 6 mg of AC assisted TiO2@AC
Table 3
Comparison of photocatalytic degradation of Rh–B by different photocatalysts.
Photocatalyst Synthesis Method Poll./ Catalyst Dosage photocatalytic efficiency Time of degradation Source of irradiated Ref.
dye (mg/L) (%) (min) light
AC on TiO2 was investigated and the results demonstrated that there was Declaration of competing interest
significant reduction in band gap of TiO2 with the addition of AC was
achieved up to 17.17 % and 34.57 % in direct and indirect band gap The authors declare that they have no known competing financial
measurements respectively. This study paves the path for the develop- interests or personal relationships that could have appeared to influence
ment of efficient photocatalytic materials for the elimination of organic the work reported in this paper.
pollutants. It is important to recognize that the findings of this study
have significant implications for the development of new photocatalytic Appendix A. Supplementary data
materials that can be used for the elimination of dyes from industrial
wastewater. Supplementary data to this article can be found online at https://doi.
org/10.1016/j.ceramint.2024.07.440.
CRediT authorship contribution statement
References
Syed Nizam uddin Shah Bukhari: Writing – original draft, Visual-
ization, Methodology, Investigation, Data curation, Conceptualization. [1] N.N. Zulu, Water scarcity and household food security: a case of ulundi local
municipality in KwaZulu-natal, South Africa, in: Handbook of Research on
Aqeel Ahmed Shah: Writing – review & editing, Validation, Supervi- Resource Management and the Struggle for Water Sustainability in Africa, IGI
sion, Conceptualization. Wen Liu: Writing – review & editing, Super- Global, 2022, pp. 127–148.
vision, Resources, Project administration, Conceptualization. Iftikhar [2] D. Yaseen, M. Scholz, Technology Textile dye wastewater characteristics and
constituents of synthetic effluents, A Critical Rev. 16 (2019) 1193–1226.
Ahmed Channa: Software, Investigation, Formal analysis, Data cura- [3] F. Sanakousar, C. Vidyasagar, V. Jiménez-Pérez, et al., Recent progress on visible-
tion. Ali Dad Chandio: Validation, Software, Resources, Methodology, light-driven metal and non-metal doped ZnO nanostructures for photocatalytic
Data curation. Imran Ali Chandio: Software, Investigation, Formal degradation of organic pollutants 140 (2022) 106390.
[4] W. Shi, M. Li, X. Huang, et al., Construction of CuBi2O4/Bi2MoO6 pn heterojunction
analysis, Data curation. Zafar Hussain Ibupoto: Writing – review & with nanosheets-on-microrods structure for improved photocatalytic activity
editing, Supervision, Project administration, Data curation. towards broad-spectrum antibiotics degradation 394 (2020) 125009.
41296
S.N.U.S. Bukhari et al. Ceramics International 50 (2024) 41285–41298
[5] M. Beija, C.A. Afonso, J. Martinho, Synthesis and applications of Rhodamine [37] Jetani G. Habibi, M. Rahmani, TiO2/GO nanocomposites: synthesis,
derivatives as fluorescent probes 38 (2009) 2410–2433. characterization, and DSSC application 135 (2020) 720.
[6] Shengnan Xu, Gang Xiao, et al., A reusable chitosan/TiO2@ g-C3N4 nanocomposite [38] H. Lu, Y. Zhou, S. Vongehr, et al., Effects of hydrothermal temperature on
membrane for photocatalytic removal of multiple toxic water pollutants under formation and decoloration characteristics of anatase TiO2 nanoparticles 55 (2012)
visible light, Water Sci. Technol. 83 (12) (2021) 3063–3074. 894–902.
[7] R. Nithya, C. Sivasankari, A. Thirunavukkarasu, Electronic waste generation, [39] S. Dessai, M. Ayyanar, S. Amalraj, et al., Bioflavonoid mediated synthesis of TiO2
regulation and metal recovery, A Rev. 19 (2021) 1347–1368. nanoparticles: characterization and their biomedical applications 311 (2022)
[8] Mohd Khairul Hanif Mohd Nazri, Norzahir Sapawe, A short review on 131639.
photocatalytic toward dye degradation, Mater. Today: Proc. 31 (2020) A42–A47. [40] A. Jimoh, B. Akpeji, S. Azeez, et al., Biosynthesis of Ag and TiO2 nanoparticles and
[9] H. Zangeneh, A. Zinatizadeh, M. Habibi, et al., Photocatalytic oxidation of organic the evaluation of their antibacterial activities 141 (2022) 109503.
dyes and pollutants in wastewater using different modified titanium dioxides, [41] M.K. Singh, M.S.J.O. Mehata, Phase-dependent optical and photocatalytic
A Comparative Rev. 26 (2015) 1–36. performance of synthesized titanium dioxide (TiO2) nanoparticles 193 (2019)
[10] W. Fu, G. Li, Y. Wang, et al., Facile formation of mesoporous structured mixed- 163011.
phase (anatase/rutile) TiO2 with enhanced visible light photocatalytic activity 54 [42] S. Singh, I.C. Maurya, A. Tiwari, et al., Green synthesis of TiO2 nanoparticles using
(2018) 58–61. Citrus limon juice extract as a bio-capping agent for enhanced performance of dye-
[11] Diana Rakhmawaty Eddy, Muhamad Diki Permana, et al., Heterophase polymorph sensitized solar cells 28 (2022) 101652.
of TiO2 (Anatase, Rutile, Brookite, TiO2 (B)) for efficient photocatalyst: fabrication [43] Z. Razzaq, A. Khalid, P. Ahmad, et al., Photocatalytic and antibacterial potency of
and activity, Nanomaterials 13 (4) (2023) 704. titanium dioxide nanoparticles: a cost-effective and environmentally friendly
[12] A. Khalid, P. Ahmad, A.I. Alharth, et al., A practical method for incorporation of Fe media for treatment of air and wastewater 11 (2021) 709.
(III) in Titania matrix for photocatalytic applications 8 (2021) 045006. [44] Z. Li, J. Li, Y. Wang, et al., Synthesis and application of surface-imprinted activated
[13] A. Khalid, P. Ahmad, A.I. Alharthi, et al., Synergistic effects of Cu-doped ZnO carbon sorbent for solid-phase extraction and determination of copper (II) 117
nanoantibiotic against Gram-positive bacterial strains 16 (2021) e0251082. (2014) 422–427.
[14] B. Mondol, A. Sarker, A. Shareque, et al., Preparation of activated carbon/TiO2 [45] B.D. Ossonon, D.J.R.A. Bélanger, Synthesis and characterization of sulfophenyl-
nanohybrids for photodegradation of reactive red-35 dye using sunlight 1 (2021) functionalized reduced graphene oxide sheets 7 (2017) 27224–27234.
54–66. [46] Q. He, Z. Hu, Y. Jiang, et al., Preconcentration of Cu (II), Fe (III) and Pb (II) with 2-
[15] X. Liu, Y. Li, D. Deng, et al., A one-step nonaqueous sol–gel route to mixed-phase ((2-aminoethylamino) methyl) phenol-functionalized activated carbon followed by
TiO2 with enhanced photocatalytic degradation of Rhodamine B under visible light ICP-OES determination 175 (2010) 710–714.
18 (2016) 1964–1975. [47] E. Alsharaeh, T. Bora, A. Soliman, et al., Sol-gel-assisted microwave-derived
[16] D.A. Hanaor, C. Sorrell, Review of the anatase to rutile phase transformation 46 synthesis of anatase Ag/TiO2/GO nanohybrids toward efficient visible light phenol
(2011) 855–874. degradation 7 (2017) 133.
[17] J. Hu, S. Zhang, Y. Cao, et al., Novel highly active anatase/rutile TiO2 [48] Dagmar Samešová, et al., Determination of aerobic and anaerobic biological
photocatalyst with hydrogenated heterophase interface structures for degradability of waste tires, Environ. Sci. Eur. 35 (1) (2023) 19.
photoelectrochemical water splitting into hydrogen 6 (2018) 10823–10832. [49] P. Kongsong, L. Sikong, S. Niyomwas, et al., Photocatalytic antibacterial
[18] Meltem Asiltürk, Şadiye Şener, TiO2-activated carbon photocatalysts: preparation, performance of glass fibers thin film coated with N-doped SnO 2/TiO 2 2014
characterization and photocatalytic activities, Chem. Eng. J. 180 (2012) 354–363. (2014).
[19] S. Sudhagar, S.S. Kumar, I.I. Premkumar, et al., UV-and visible-light-driven TiO2/ [50] V. Jain, M.C. Biesinger, M. Linford, The Gaussian-Lorentzian Sum, Product, and
La2O3 and TiO2/Al2O3 nanocatalysts: synthesis and enhanced photocatalytic Convolution (Voigt) functions in the context of peak fitting X-ray photoelectron
activity 128 (2022) 282. spectroscopy (XPS) narrow scans 447 (2018) 548–553.
[20] X. Wang, H. Cheng, G. Ye, et al., Key factors and primary modification methods of [51] D.N.G. Krishna, J. Philip, Review on surface-characterization applications of X-ray
activated carbon and their application in adsorption of carbon-based gases, A Rev. photoelectron spectroscopy (XPS): Recent developments and challenges 12 (2022)
287 (2022) 131995. 100332.
[21] Z.M. Yunus, A. Al-Gheethi, N. Othman, et al., Advanced methods for activated [52] J. Shao, W. Sheng, M. Wang, et al., In situ synthesis of carbon-doped TiO2 single-
carbon from agriculture wastes, A Comprehensive Rev. 102 (2022) 134–158. crystal nanorods with a remarkably photocatalytic efficiency 209 (2017) 311–319.
[22] R. Bhattacharya, A review on production and application of activated carbon from [53] Z. Xiong, X. Zhao, Nitrogen-doped titanate-anatase core–shell nanobelts with
discarded plastics in the context of ‘waste treats waste’ 325 (2023) 116613. exposed {101} anatase facets and enhanced visible light photocatalytic activity
[23] H. Karimi-Maleh, C. Karaman, O. Karaman, et al., Nanochemistry approach for the 134 (2012) 5754–5757.
fabrication of Fe and N co-decorated biomass-derived activated carbon [54] J. Liu, Q. Zhang, J. Yang, et al., Facile synthesis of carbon-doped mesoporous
frameworks: a promising oxygen reduction reaction electrocatalyst in neutral anatase TiO2 for the enhanced visible-light driven photocatalysis 50 (2014)
media 12 (2022) 429–439. 13971–13974.
[24] H. Susanto, R. Mulyanti, T. Istirokhatun, et al., Treatment of saline domestic [55] P. Shao, J. Tian, Z. Zhao, et al., Amorphous TiO2 doped with carbon for visible light
wastewater using nanofiltration membrane coupled with activated carbon photodegradation of rhodamine B and 4-chlorophenol 324 (2015) 35–43.
adsorption 19 (2022) 62–73. [56] M.A. Mohamed, W.N.W. Salleh, J. Jaafar, et al., Carbon as amorphous shell and
[25] M. Salt, S.T.S. Yuen, N. Ashwath, et al., "Solid waste landfilling: concepts, interstitial dopant in mesoporous rutile TiO2: Bio-template assisted sol-gel
Processes, Technologies (2018) 677–688. synthesis and photocatalytic activity 393 (2017) 46–59.
[26] M.A. Bhatti, K.F. Almaani, A.A. Shah, et al., Low temperature aqueous chemical [57] A.M. Abdullah, N.J. Al-Thani, K. Tawbi, et al., Carbon/nitrogen-doped TiO2: New
growth method for the doping of W into ZnO nanostructures and their synthesis route, characterization and application for phenol degradation 9 (2016)
photocatalytic role in the degradration of methylene blue (2022) 1–12. 229–237.
[27] S. Samuchiwal, A. Bhattacharya, A. Malik, Treatment of textile effluent using an [58] Y.-F. Lee, K.-H. Chang, C.-C. Hu, et al., Synthesis of activated carbon-surrounded
anaerobic reactor integrated with activated carbon and ultrafiltration unit (AN- and carbon-doped anatase TiO2 nanocomposites 20 (2010) 5682–5688.
ACF-UF process) targeting salt recovery and its reusability potential in the pad- [59] B.E. Azar, A. Ramazani, S.T. Fardood, et al., Green synthesis and characterization
batch process 40 (2021) 101770. of ZnAl2O4@ ZnO nanocomposite and its environmental applications in rapid dye
[28] M. Aravind, M. Amalanathan, M.S.M. Mary, et al., Enhanced photocatalytic and degradation 208 (2020) 164129.
biological observations of green synthesized activated carbon, activated carbon [60] H. Liu, C. Wang, G. Wang, Photocatalytic advanced oxidation processes for water
doped silver and activated carbon/silver/titanium dioxide nanocomposites 32 treatment: recent advances and perspective 15 (2020) 3239–3253.
(2022) 267–279. [61] A. Habibi-Yangjeh, S. Asadzadeh-Khaneghah, S. Feizpoor, et al., Review on
[29] G. Murali, S. Gopalakrishnan, S.K. Lakhera, et al., Enhanced solar light-driven heterogeneous photocatalytic disinfection of waterborne, airborne, and foodborne
hydrogen evolution of activated carbon sphere supported TiO2 hybrid viruses: can we win against pathogenic viruses? 580 (2020) 503–514.
nanocomposites 128 (2022) 109226. [62] D. Zhu, Q.J.E.N. Zhou, Monitoring, Management. Action and mechanism of
[30] F. Scarpelli, T.F. Mastropietro, T. Poerio, et al., Mesoporous TiO2 thin films: State semiconductor photocatalysis on degradation of organic pollutants in water
of the art 508 (2018) 135–142. treatment, Review 12 (2019) 100255.
[31] A.C. Martins, A.L. Cazetta, O. Pezoti, et al., Sol-gel synthesis of new TiO2/activated [63] Z. Yigit, H.J.W. Inan, Air, S.P. Focus, A study of the photocatalytic oxidation of
carbon photocatalyst and its application for degradation of tetracycline 43 (2017) humic acid on anatase and mixed-phase anatase–rutile TiO2 nanoparticles 9 (2009)
4411–4418. 237–243.
[32] B. Xing, C. Shi, C. Zhang, et al., Preparation of TiO2/activated carbon composites [64] G. Naresh, V. Kumar, B. Sasikumar, et al., Improved H2 yields over Cu-Ni-TiO2
for photocatalytic degradation of RhB under UV light irradiation 2016 (2016), 3-3. under solar light irradiation: Behaviour of alloy nano particles on photocatalytic
[33] K. Natarajan, T.S. Natarajan, H.C. Bajaj, et al., Rutile phase dominant TiO2 formed H2O splitting 299 (2021) 120654.
by thermal treatment and its high photocatalytic activity under narrow spectrum [65] M.A. Bhatti, S.J. Gilani, A.A. Shah, et al., Effective removal of methylene blue by
ultraviolet light emitting diodes 6 (2018) 015049. surface alteration of TiO2 with Ficus Carica leaf extract under visible light 12
[34] D. Ziental, B. Czarczynska-Goslinska, D.T. Mlynarczyk, et al., Titanium dioxide (2022) 2766.
nanoparticles: prospects and applications in medicine 10 (2020) 387. [66] S. Sun, Y. Xue, D. Yang, et al., Bismuth pyrochlores with varying Fe/Co ratio for
[35] O.O. Apeh, U.K. Chime, S. Agbo, et al., Properties of nanostructured ZnO thin films efficient Multi-functional Catalysis: Structure evolution versus Photo-and Electro-
synthesized using a modified aqueous chemical growth method 6 (2019) 056406. catalytic activities 448 (2022) 137580.
[36] Y.E. Du, Y. Bai, Y. Liu, et al., One-Pot Synthesis of [111]-/{010} Facets Coexisting [67] G. Nabi, N. Khalid, M.B. Tahir, et al., A review on novel eco-friendly green
Anatase Nanocrystals with Enhanced Dye-Sensitized Solar Cell Performance 1 approach to synthesis TiO2 nanoparticles using different extracts 28 (2018)
(2016) 6632–6640. 1552–1564.
41297
S.N.U.S. Bukhari et al. Ceramics International 50 (2024) 41285–41298
[68] C.J.E.N. Thambiliyagodage, Monitoring, management, Efficient photocatalysis of [81] Emmanuel Nyankson, Nathaniel Yeboah, et al., The effect of synthesis route on the
carbon coupled TiO2 to degrade pollutants in wastewater–A review 18 (2022) photocatalytic performance of Ag-TiO2 using rhodamine b dyes, pesticides, and
100737. pharmaceutical waste as model pollutants, Mater. Res. Express 9 (9) (2022)
[69] J. Li, W. Ma, C. Chen, et al., Photodegradation of dye pollutants on one- 094001.
dimensional TiO2 nanoparticles under UV and visible irradiation 261 (2007) [82] Shijie Tong, Jin Zhou, Ling Ding, et al., Preparation of carbon quantum dots/TiO2
131–138. composite and application for enhanced photodegradation of rhodamine B,
[70] F. Moradnia, S.T. Fardood, A. Ramazani, et al., Magnetic Mg0. 5Zn0. 5FeMnO4 Colloids Surf. A Physicochem. Eng. Asp. 648 (2022) 129342.
nanoparticles: green sol-gel synthesis, characterization, and photocatalytic [83] Vitória De Lima Brombilla, et al., Biochar derived from yerba-mate (Ilex
applications 288 (2021) 125632. paraguariensis) as an alternative TiO2 support for enhancement of photocatalytic
[71] T. Luttrell, S. Halpegamage, J. Tao, et al., Why is anatase a better photocatalyst activity toward Rhodamine-B degradation in water, Chem. Eng. Commun. 209 (10)
than rutile?-Model studies on epitaxial TiO2 films 4 (2014) 4043. (2022) 1334–1347.
[72] G. Li, K. Fang, Y. Ou, et al., Surface study of the reconstructed anatase TiO2 (001) [84] Ntakadzeni Madima, et al., Fabrication of magnetic recoverable Fe3O4/TiO2
surface 31 (2021) 1–13. heterostructure for photocatalytic degradation of rhodamine B dye, Inorg. Chem.
[73] P. Makuła, M. Pacia, W. Macyk, in: How to Correctly Determine the Band Gap Commun. 145 (2022) 109966.
Energy of Modified Semiconductor Photocatalysts Based on UV–Vis Spectra, vol. 9, [85] Y. Luo, Z. Zhang, S. Chen, A. Jiang, X. Chen, X. Zhu, Y. Di, Activation of
ACS Publications, 2018, pp. 6814–6817. peroxymonosulfate by halloysite nanotube-supported TiO2 composite for efficient
[74] Mohammad Saleh Shafeeyan, et al., A review on surface modification of activated photocatalytic degradation of Rhodamine B, Opt. Mater. 147 (2024) 114669.
carbon for carbon dioxide adsorption, J. Anal. Appl. Pyrol. 89 (2) (2010) 143–151. [86] D. Xu, P. Wang, B. Shen, Development of tio2-reduced graphene oxide
[75] V.L. Chandraboss, J. Kamalakkannan, S. Senthilvelan, Synthesis of activated nanocomposites and their enhanced photocatalytic and photovoltaic performance,
charcoal supported Bi-doped TiO2 nanocomposite under solar light irradiation for Dig. J. Nanomater. Biostruct. 11 (1) (2016).
enhanced photocatalytic activity, Appl. Surf. Sci. 387 (2016) 944–956. [87] Rong He, Wei He, Ultrasonic assisted synthesis of TiO2–reduced graphene oxide
[76] M.N. Rashed, M. Eltaher, A. Abdou, Adsorption and photocatalysis for methyl nanocomposites with superior photovoltaic and photocatalytic activities, Ceram.
orange and Cd removal from wastewater using TiO2/sewage sludge-based Int. 42 (5) (2016) 5766–5771.
activated carbon nanocomposites 4 (2017) 170834. [88] Salam Hussein Alwan, et al., Visible light-driven photocatalytic degradation of
[77] Xiaobing Yang, et al., MIL-125 (Ti) derived Ag doped tablet-like TiO2@ carbon Rhodamine B dye onto TiO2/rGO nanocomposites, Mater. Today Commun. 33
composites as efficient photocatalyst for degradation of rhodamine B, J. Solid State (2022) 104558.
Chem. 320 (2023) 123862. [89] Mohamed Nageeb Rashed, et al., Photocatalytic degradation of Rhodamine-B dye
[78] Jong-Hoon Lee, Seong-Jun Mun, et al., Promoted charge separation and specific using composite prepared from drinking water treatment sludge and nano TiO2,
surface area via interlacing of N-doped titanium dioxide nanotubes on carbon Environ. Qual. Manag. 31 (3) (2022) 175–185.
nitride nanosheets for photocatalytic degradation of Rhodamine B, Nanotechnol. [90] Zaid H. Mahmoud, et al., The efficacy of samarium loaded titanium dioxide (Sm:
Rev. 11 (1) (2022) 1592–1605. TiO2) for enhanced photocatalytic removal of rhodamine B dye in natural sunlight
[79] Rabeya Sultana, Samia Islam Liba, et al., Enhanced photocatalytic activity in Rh-B exposure, J. Mol. Struct. 1253 (2022) 132267.
dye degradation by Mn and B co-doped mixed phase TiO2 photocatalyst under [91] Z. Chen, S. Yu, et al., F co-doping Ag/TiO2 with visible light photocatalytic
visible light irradiation, Surface. Interfac. 42 (2023) 103302. performance toward degrading Rhodamine B, Environ. Res. 232 (2023) 116311.
[80] Mohadeseh Elhami, et al., Process efficiency and life cycle assessment of novel
waste derived Chl/TiO2 photocatalyst for rhodamine B removal, J. Water Process
Eng. 56 (2023) 104425.
41298