1. Introduction
Water is an essential resource for the growth of the natural environment on the earth and people. Water pollution with organic dyes, heavy metals, and other contaminants (i.e, pesticides, steroid hormones, antibiotics) is a current environmental issue [
1]. Synthetic organic dyes are omnipresent in many application areas of the textile, tannery, cosmetic and food industries, and medicine [
1]. The dying process in the textile and tannery industries can release a vast majority of organic dyes due to the nature low efficiency of the dying process with approximately 15% of dye lost [
1,
2]. Organic dye pollution in the aquatic environment poses a threat to animal or human health and causes adverse effects on aquatic biota and water ecosystems [
1,
3]. Semiconductor photocatalysis for environmental applications was introduced [
4], and the field has been greatly developed over the years [
5,
6].
Titanium dioxide (TiO
2) is well-known to be the most practical and prevalent photocatalyst owing to its high photocatalytic activity, cost-effectiveness, chemical stability, abundant, and nontoxic [
5]. Commercial powder of TiO
2 Degussa P25 (a flame-made multiphasic TiO
2 nanoparticles) was developed, and it has been widely used as a benchmark material for studying photocatalytic mechanisms, materials, and processes [
7]. TiO
2 P25 contains anatase, rutile, and a small amount of amorphous phase, which were reported to be 77.1% anatase, 15.9% rutile, and 7.0% amorphous TiO
2 [
7]. However, TiO
2 is a wide band gap semiconductor with E
g = 3.2 eV for anatase and 3.0 for rutile [
8], and consequently, TiO
2 only works in the ultraviolet region, which accounts for less than 5% of the total energy of the solar spectrum [
9]. A drawback of TiO
2 nanoparticles is the fast recombination of photoexcited carriers (electrons and holes) [
10], and low specific surface area that limits the contaminant adsorption capability and the photocatalytic activity.
TiO
2 P25 (a powder form photocatalyst) offers relatively high photocatalytic degradation of organic dyes [
11] and pharmaceuticals [
11,
12,
13,
14] and a high possibility toward practical scale application. However, TiO
2 P25 use in the form of suspension (slurry) may pose an ecological risk to aquatic organisms [
15], which should need an expensive filtration process to separate the suspension catalyst from the treated water. Therefore, many nanostructured TiO
2 films developed on rigid substrates (e.g. TiO
2 nanotubes [
16], TiO
2 nanowires on TiO
2 nanotube arrays [
17,
18,
19]) have been studied for photocatalyst, solar energy conversion, and other applications [
16]. The film forms of TiO
2 nanomaterials face a big challenge in scaling up to the practical scale of water treatment. Another approach is that TiO
2 nanomaterial is loaded on/in another bigger nano/micro-scale structure to form a heterostructure or a composite. Heterostructure and composite photocatalysts can offer activity enhancement due to some mechanism related to manipulating the photoresponse region and the fate of the electron/hole pairs [
9]. In this way, TiO
2 has been impregnated on several porous supports such as silica, alumina, zeolite, and carbon materials [e.g., activated carbon (AC)] [
20,
21,
22]. TiO
2/AC has been found to possess higher photocatalytic activity than bare TiO
2 as it combines the good photocatalytic activity of TiO
2 and the high surface area of AC [
21,
22,
23,
24,
25,
26,
27,
28]. In addition, TiO
2 nanoparticles loaded in the AC immobilization can resolve the limitation of using TiO
2 suspension, while AC is a very economical–useful adsorbent, and it is characterized by high surface area, micro- to macro-pore structures, and a high degree of surface reactivity [
29].
Many studies focused on developing TiO
2/AC with structural, compositional, and surface functional modification [
21,
22,
23,
24,
25,
26,
27,
28], which usually needs the use of chemicals and laboratory equipment. Noticeably, a complicated preparation process of TiO
2/AC can hinder the practical application at the household level, where non-specialized people will the the operator of their household water treatment tank or plant. In this study, we prepared TiO
2 P25/AC using the two commercial raw materials by a facile and low-cost method by mixing TiO
2/AC mechanically at different mass ratios. Differ from the previous studies, the compositional weight portions of TiO
2/AC were in some specific range such as TiO
2/ (1–15 wt.%)-AC [
28], TiO
2/ (5–75 wt.%)-AC prepared by sol-gel method [
21], while the weight ratios of TiO
2/AC in this study were TiO
2/AC = 5:0, TiO
2/AC = 4:1, TiO
2/AC = 3:2, TiO
2/AC = 2:3, TiO
2/AC = 1:4, TiO
2/AC = 0:5. It is worthy to mention that the use mass ratio unit for the composite should be easier and more convenient for normal people to apply in practice when they implement TiO
2/AC material into their household water treatment plant. Moreover, this study has two reference samples (i.e., TiO
2, AC) to get insight into the role of either TiO
2 or AC in the photocatalytic degradation of methylene blue, which is a typical organic dye and considered a good model for studying photocatalyst activity and process. This study provides the effect of mixing weight ratios of TiO
2/AC on the morphological, structural, compositional, and photocatalytic properties of the composite.
3. Results and Discussion
Figure 2a presents the XRD pattern of TiO
2 P25 nanoparticles (S1), TiO
2/AC composites prepared at various mass ratios (S2 – S5), and AC (S6). Generally, TiO
2 had dominant anatase phase characterized by diffraction peaks of A(101) at 25.3°, A(004) at 37.9°, A(200) at 48.2°, A(204) at 62.8°, and minor rutile phase with the observed peaks of R(110) at 27.5° and R(211) at 54.1°. Meanwhile, the AC exhibited the diffraction peaks of graphite [e.g., G(002) at 26.6°, G(020) at 45.5°], carbon C(100) at 20.9°, and 3D carbon structures with peaks G
3D(002) at 30.0°. The intensity of TiO
2 and AC peaks varied reasonably with the evolution of TiO
2/AC content (see
Figure 2a). The dominant anatase phase over the rutile one for TiO
2 Degussa P25 in this study is consistent well with results reported in ref. [
30].
The crystallite size (
D) in TiO
2 P25 and AC were estimated using the Scherrer equation and TiO
2 (101) and G(200) peaks, respectively. The Scherrer equation is given as,
D = 0.9
λ/βcos
θ, where
λ,
β, and
θ are the X-ray wavelength, full width at half maximum of the diffraction peak, and Bragg diffraction angle, respectively [
19,
21,
31]. The
D value for TiO
2 P25 in S1–S5 samples was a range of 20.6 – 21.0 nm, while the
D of AC was 48.9 nm. The
D of TiO
2 in this study is comparable with that of pristine TiO
2 synthesized by the sol-gel method (D = 17.5 nm), smaller than the D values of 43.4 – 109.0nm for the TiO
2 annealed thermally at 200 – 500°C for 120 min [
26]. Noticeably, the crystallite size (48.9 nm) for the present AC is very close to the
D of 50 nm for the commercial AC used in ref. [
26].
Figure 3 shows the surface morphology of the studied materials. TiO
2 P25 exhibited uniform spherical nanoparticles (NPs) with a size of 25 – 35 nm, while AC was micron- and sub-micron particles. For TiO
2/AC composites, TiO
2 nanoparticles were well dispersed and decorated tightly on AC particles (
Figure 3). The SEM images also reflect the contents of TiO
2 and AC in the composites, e.g., TiO
2 content decreases while AC content increases when we observe the SEM images from S2 to S5 (
Figure 3). The elemental composition and distribution were illustrated by the EDS result (S4) in
Figure 4. This composite has 18.04 at.% C, 52.18 at.% O, 19.00 at.% Ti, which indicates a composition of AC/TiO
2. Notably, small contents of Na (5.5 at.%), Si (5.25 at.%), and Pt (0.03 at.%) were observed due to the remaining after the surface activation process for P25 using NaOH 5 M, the Si substrate, and Pt coating for taking SEM images, respectively. In addition, the EDS mapping in
Figure 4 suggests a uniform elemental distribution, uniform decoration of TiO
2 on AC, and possibly loading of TiO
2 NPs inside the micro-pores and micro-channels of AC. The SEM and EDS results indicate the tight binding between TiO
2 and AC that should support the carrier transport for enhancing photocatalytic activity.
To gain insight into the crystalline structure of TiO
2/AC composites, a typical Raman spectrum of TiO
2/AC was examined. In
Figure 5, the spectrum exhibited characteristic peaks of TiO
2 predominant anatase phase at 146 cm
-1 (E
g(1)), 200 cm
-1 (E
g(2)), 396 cm
-1 (B
1g), 516 cm
-1 (A
1g), and 635 cm
-1 (E
g) as well as two graphite carbon peaks at 1327 cm
-1 (D-band) and 1588 cm
-1 (G-band) (
Figure 5). The D-band is associated with asymmetric lattices and bond-angle disorders in graphitic structures [
25], while G-band comes from the doubly-degenerate iTO and LO phonon with E
2g symmetry at the Brillouin zone center [
32]. This Raman results for TiO
2/AC agreed well with the Raman results for the TiO
2/AC in ref. [
25].
The photocatalytic activities of TiO
2, AC, and TiO
2/AC with different mixing mass ratios were studied by monitoring the photodegradation kinetics of MB (an organic substance, and popular use as an organic pollutant model) under UV-Vis irradiation (120 mW/cm
2).
Figure 6a shows the evolution of the MB absorption peak (at ~655 nm) over the photocatalytic reaction time (
t) using the TiO
2/AC (S2). The peak intensity (associated with MB concentration) decreased with
t, and this behavior is true for the other photocatalysts in this study (S1, S3, S4, S5, S6). The photodegradation kinetics by photolysis and photocatalytic processes using the materials are shown in
Figure 6b, which obeys the Langmuir–Hinshelwood kinetic model with the first-order reaction rate constant (
k), C
t = C
o×e
-kt, where C
t is the concentration of MB at time t (mg/L), C
o is the initial MB concentration (mg/L).
The solid lines in
Figure 6b are the fitting curves using the kinetic model. The fittings yield the
k values of the photolysis (P) and photocatalytic processes using the S1–S6 materials (see
Figure 6c). The
k of the photolysis reaction was a small value of 1.8×10
-3min
-1, while the
k values were 32.1×10
-3 min
-1 (S1), 55.2×10
-3 min
-1 (S2), 38.5×10
-3 min
-1 (S3), 37.5×10
-3 min
-1 (S4), 28.9×10
-3 min
-1 (S5), 18.7×10
-3 min
-1 (S6). This indicates that MB degradation is much faster and more effective by using the photocatalysts as compared to the photolysis process. In addition, the composites of TiO
2/AC (S2, S3, S4) exhibited an enhancement in the photocatalyst activity over either the pristine TiO
2 (S1) or AC (S6). Among the investigated TiO
2/AC composites with various mass mixing ratios (S2 – S5), the S2 possessed the highest photocatalyst activity, suggesting that TiO
2/AC = 4:1 is the optimal mass mixing ratio between TiO
2 P25 and micron- and sub-micron AC. Further increased AC and decreased TiO
2 contents to TiO
2/AC = 3:2 and TiO
2/AC = 2:3 also exhibited higher photocatalytic activities than that for TiO
2. Meanwhile, the sufficient low TiO
2 and high AC contents for the TiO
2/AC = 1:4 (S5) lead to the decrease of activity as compared to pristine TiO
2 (see the dashed line in
Figure 6c). Noticeably, the observed decrease in MB concentration over
t for AC (S6) material is attributed to the excellent adsorption characteristic of porous carbon materials (
Figure 6c inset) [
27,
33,
34].
AC/TiO
2 composites with mass mixing ratios of (4:1), (3:2), and (2:3) have higher MB removal performance than either TiO
2 or AC, which is attributed to the synergistic effect of the high adsorption capability of AC and the high photocatalytic activity of TiO
2. As illustrated in
Figure 6d, under UV-VIS irradiation, electron/hole (e
-/h
+) pairs are generated, which lead subsequently to oxidation and reduction reactions in the treated solution to generate highly active free radicals, primarily
•OH and
•O
2 ⁻ (
Figure 6d) [
27,
33,
34], which in turn degrade MB dye. It is well-known that the recombination rate of e
-/h
+ in TiO
2 is fast. Meanwhile, in TiO
2/AC composites, the photogenerated carriers can be transferred between TiO
2 and AC, allowing to increase in e
-/h
+ separation and reducing the e
-/h
+ recombination rate, and consequently enhancing the photocatalytic activity. The AC (S6) removes MB with
k of 18.7×10
-3 min
-1 (58.3% of
k for TiO
2), suggesting the AC has high adsorption capacity with many sufficient active adsorption sites and a large surface area [
27,
33,
34]. The lower
k value of S5 than S1 indicates that a composite with too little TiO
2 and too much AC contents (TiO
2/AC = 1:4 for the present case) will not give an enhancement of the photocatalytic activity. An advantage of a combination of a good photocatalyst with a good adsorption material (i.e., TiO
2/AC) is that TiO
2 degrades MB to create renewable active adsorption sites on the AC to adsorb more MB molecules as evidenced by the higher adsorption efficiency of pollutants near TiO
2 positions on the composite surfaces [
27,
35]. Also, TiO
2 plays the role of photocatalytic regeneration of AC, and this allows minimizing operational costs and product waste. The optimal TiO
2/AC (S2) obtained a high MB removal efficiency of 96.6% after 60 min treatment at the initial MB concentration of 10 mg/L and a composite dosage of 333.3 mg/L. This efficiency is comparable to that for the TiO
2/AC (i.e., 86.5%, 97.1%, and 99.4%, depending on the type of AC) under similar experimental conditions (i.e., reaction time of 60 min, UV light source, catalyst dosage of 400 mg/L, and MB initial concentration of 20 mg/L) [
27].