Next Article in Journal
Response Surface Methodology Approach for Optimized Biodiesel Production from Waste Chicken Fat Oil
Previous Article in Journal
Tandem Catalysis: Synthesis of Nitrogen-Containing Heterocycles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Boron-Doped TiO2-CNT Nanocomposites with Improved Photocatalytic Efficiency toward Photodegradation of Toluene Gas and Photo-Inactivation of Escherichia coli

1
Department of Materials Science and Engineering, University of Seoul, Seoul 02504, Korea
2
Department of Physics, National Dong Hwa University Shou-Feng, Hualien 97401, Taiwan
3
Department of Chemistry, Shivaji University, Kolhapur 416004, India
4
Department of Chemistry, D.B.F. Dayanand College of Arts and Science, Solapur 413002, India
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(6), 632; https://doi.org/10.3390/catal10060632
Submission received: 14 May 2020 / Accepted: 3 June 2020 / Published: 5 June 2020
(This article belongs to the Section Photocatalysis)

Abstract

:
An in-situ sol-gel method was used for the synthesis of boron-doped TiO2-CNT nanocomposites with varied boron concentrations from 1 to 4 mol%. The synthesized nanocomposites were characterized by various techniques, namely XRD, UV-DRS, TEM, PL, and XPS; all results show that 3 mol% B-TiO2-CNT nanocomposites have superior properties to pure TiO2, 3B-TiO2 nanoparticles, and other nanocomposites. TEM images clearly show the B-TiO2 nanoparticles decorated on the CNT surface. Photo-luminescence studies confirm that boron doping of up to 3 mol% in TiO2-CNT nanocomposites reduces the electron-hole pair recombination rate. The photocatalytic performance of the B-TiO2-CNT nanocomposites was tested against the photodegradation of toluene gas and the photocatalytic inactivation of E. coli in the presence of UV and visible light respectively. B-TiO2-CNT (3 mol%) nanocomposites show the highest photocatalytic activity.

1. Introduction

The nonstop rise in environmental pollution is responsible for several critical and detrimental global problems. For example, the release of large amounts of toxic dyes into wastewater, as well as the release of toxic environmental gases such as volatile organic compounds (VOCs), are destroying our ecosystem. Another worldwide problem arises from continuous and serious health issues due to infections from microorganisms. Moreover, indoor air quality is increasingly deteriorating because various industries are releasing unpurified VOCs into the environment [1]. It is well known that VOCs are highly toxic and hazardous to human health. This category of compounds includes toluene, xylene, ethylbenzene, and benzene, of which toluene is the most generally known [2,3]. The inhalation of toluene can harm the central nervous system, causing loss of hearing and color vision, as well as the loss of cognitive abilities and muscle memory [4,5]. Therefore, it is necessary to develop effective technology to degrade and eliminate VOCs from the environment. In the literature, the photocatalytic degradation of VOCs has recently attracted considerable attention, as it is considered to be one of the most potent methods for environmental purification [6,7].
Infectious diseases that are caused by pathogenic microorganisms may result in lifelong damage if detection and medical treatment are not managed in a timely fashion [8]. and even death of a patient can occur [9]. Pathogenic bacteria, including Staphylococcus aureus, Salmonella enterica, Streptococcus pyogenes, Streptococcus pneumonia and Escherichia coli, cause serious diseases [10,11,12,13,14] and enjoy an environmentally stable and high production rate. To control these types of infections, many strategies are available, such as heat treatment, UV light irradiation, chemical agents, and a few biological agents, but these strategies are expensive and produce toxic byproducts in the environment.
Photocatalytic inactivation of bacteria has attracted great attention [15]. In particular, TiO2 nanoparticles (NPs) are attractive photocatalysts due to features such as corrosion resistance, low cost, structure, thermal stability, and environmental friendliness [16,17]. The large surface area of TiO2 NPs lends them to applications in various fields, such as degradation of various contaminants [16,18] and oxidative reactions [19,20]. Recently, TiO2 NPs have been applied in water splitting, sensitized solar cells, fuel cells, the paint industry, and cosmetics, as well as in biomedical applications [21,22]. Despite these advantages, TiO2 NPs suffer from some limitations: high bandgap energy pushes photocatalytic activity into the UV region of the electromagnetic spectrum, and a high electron-hole pair recombination rate limits the use of TiO2 NPs in large-scale applications [23].
To surmount these drawbacks, researchers are applying different strategies, like metal doping [24], non-metal doping [25], nanocomposites with metal oxides, carbon nanostructures [26], and noble metal loading [27]. Among these strategies, non-metal doping, using carbon, nitrogen, iodine, boron, sulfur, surface treatment [28], and fabricated nanocomposites with carbon nanotubes (CNT) attracts grate attention. CNTs have an extraordinary property such as charge transmission, electron-conducting, high mechanical strength, hollow layered structure, and large surface. Due to this, it makes an auspicious candidate as dopants and supports various photocatalysts [29]. With these advantages, CNTs also have limitations, the strong van der Waals interactions and the hydrophobicity of CNTs lead to poor dispersion in solvents and this inadequate chemical compatibility highly restricts the variety of applications [30]. To overcome this limitation there is a way to covalent surface functionalization. Surfaces functionalization improve the solubility of CNTs in various solvents to help make hybrid composites with various materials [31,32]. Such as nanocomposites between TiO2 and CNTs improved photocatalytic efficiency. Due to the formation of a heterojunction that reduces the e−h+ pair recombination, generates e−h+ pairs in presence of a photon, and helps to easy transfer of an electron to the conduction band of TiO2; CNTs play a role as an impurity by forming Ti−O−C or Ti−C defect sites that permit visible light absorption [33]. Recently, boron doping in anatase TiO2 has shown many advantages, since the presence of boron may help weaken the Ti–O bonds, making it easier to replace the oxygen with boron [34,35]. Anatase TiO2 photocatalyst doped with boron exhibits a strong absorption band in the visible region, as well as the separation of electron-hole pairs. Hong et al. [36] doped with boron in an anatase TiO2 host lattice to maximize the visible-light photocatalytic activity. Sopchenski et al. [37] prepared B-doped TiO2 and applied it to the photocatalytic inactivation of Staphylococcus aureus and Pseudomonas aeruginosa.
Besides, researchers have prepared nanocomposites of TiO2 NPs with CNTs for photodegradation of organic pollutants [38,39]. CNTs improve the optical properties of TiO2 NPs under irradiation with UV and visible light: the non-metal doping enhances the photocatalytic efficiency of TiO2 by shifting the absorption band into the visible light region, and the CNTs provide strong, stable support that reduces nanoparticle agglomeration. Therefore, the present work focuses upon a one-pot synthesis of boron-doped TiO2-CNT nanocomposites and studies their structural, morphological properties, then applies them toward the photocatalytic degradation of toluene gas and the photocatalytic inhibition of Escherichia coli (E. coli.)

2. Results and Discussion

2.1. X-ray Diffraction Studies

Figure 1 shows the X-ray diffraction (XRD) patterns of TiO2 NPs, 3B-TiO2, and B-TiO2-CNT nanocomposites. The characteristic peak at 25.3° of anatase TiO2 was observed in pure TiO2 NPs as well as in B-TiO2-CNT nanocomposites, as other peaks at 37.07°, 48.24°, 54.62°, 55.58°, 62.8°, 68.76°, 70.31°, and 75.03°. All peaks matched JCPDS card 21-1272 for anatase TiO2 [40]. Doping with boron and forming nanocomposites with CNT, increases the intensity of XRD diffraction peaks, confirming heterogeneous nucleation between the TiO2 and CNT [41]. The absence of peaks for brookite and rutile TiO2 confirmed the boron doping, and the presence of CNTs did not affect the anatase TiO2 phase. Moreover, the absence of peaks for boron or boron-containing phases confirmed that the boron dopants in TiO2 were highly dispersed [42]. In the doping process, boron replaces the oxygen in the TiO2 lattice, and in calculations at 450 °C, boron occupies interstitial positions in the TiO2 lattice. Due to the very low concentration of CNT, it not showing any peak in the XRD patterns of nanocomposites. The crystallite size of the synthesized nanomaterials was calculated by Scherrer’s formula and is listed in Table 1.

2.2. UV-Vis Diffuse Reflectance Spectroscopy

The influence of boron doping and the presence of CNT on the optical properties of TiO2 NPs is shown in Figure 2a. 3B-TiO2 and B-TiO2-CNT nanocomposite samples show higher absorbance than pure TiO2 NPs in the UV range, and also exhibit an absorption band in the visible range, above 400 nm. Moreover, the significant shift in the absorption edge of the nanocomposites towards higher wavelength confirmed a reduction in the bandgap [43]. A Tauc plot was used to determine the bandgap, as shown in Figure 2b, and values are listed in Table 1. B-TiO2 shows a redshift compared to undoped TiO2 since the p orbital of B is mixed with the 2p orbital of O [44,45,46]. The literature also reveals that non-metal doped TiO2 shows enhanced activity toward visible light by the formation of stable oxygen vacancies on the TiO2 surface [45]. Also, the presence of CNT in nanocomposites helps enhance the photocatalytic efficiency of B-TiO2. Due to the lower Fermi level of CNT as compared to TiO2 NPs, CNT may accept and store photogenerated electrons, as well as enhance the separation of electron-hole pairs [47].

2.3. Photoluminescence Analysis

Figure 3 shows the photoluminescence (PL) emission spectra of TiO2, 3B-TiO2, and B-TiO2-CNT nanocomposites. PL analysis yields important information regarding the separation and recombination of electron-hole pairs [48]. For an excitation wavelength of 285 nm, the spectra showed two emission peaks at 400 nm and 464 nm, which can be attributed to emission by free TiO2 excitons and surface Ti–OH bonds, respectively [49]. As shown in Figure 3, the TiO2, 3B-TiO2, 1B-TiO2-CNT, 2B-TiO2-CNT, and 4B-TiO2-CNT samples exhibited high-intensity peaks, while 3B-TiO2-CNT exhibited a low-intensity peak, confirming that 3B-TiO2-CNT has reduced recombination of photogenerated electron-hole pairs and high photocatalytic activity. The peak for 3B-TiO2 NPs shows lower intensity as compared to 1B-TiO2-CNT it confirms the boron doping, as well as CNT, plays an important role in reducing the recombination rate of electron-hole pair and enhancing photocatalytic activity. UV-DRS data Figure 2b shows that the excess doping of boron in TiO2 does not reduce the bandgap of TiO2. Similar phenomena were observed by other researchers and it may be due to excess doping of boron in TiO2 that could result in the aggregation of the particles which affect the optical properties of synthesized nanocomposites [43,50].

2.4. Transmission Electron Microscopy Analysis

Figure 4 shows transmission electron microscopy (TEM) images of bare TiO2, 3B-TiO2 NPs, 1B-TiO2-CNT, and 3B-TiO2-CNT nanocomposites. From Figure 4a, it is clear that the morphology of the pure TiO2 NPs is spherical, with sizes ranging from 10 to 20 nm and slight agglomeration. Figure 4c shows the same morphologies as TiO2 with slight increases in particle size due to 3 mole% boron doping TiO2, the particle sizes observed in the range from 10 to 30 nm. Figure 4e,g shows TEM images of 1B-TiO2-CNTs and 3B-TiO2-CNT respectively. Images clearly show the B-TiO2 NPs decorated on the CNT surface. confirming that CNT provide support to B-TiO2 NPs, reduces the agglomeration of nanoparticle, and increases the reactive surface area of the nanoparticles to photocatalytic activity [51]. Also, CNT are highly conductive, they permit easy movement of photogenerated electrons within the nanocomposites, improving the optical properties of TiO2 NPs [52,53]. Figure 4b,d,f,h shows lattice fringes for TiO2, 3B-TiO2 NPs, 1B-TiO2-CNT, and 3B-TiO2-CNT nanocomposites respectively. The interplanar distance for TiO2 NPs is 0.35 nm while in the 1B-TiO2-CNT is 0.36 and 3B-TiO2 0.37 nm it confirms the as the dopant concentration increasers particles size increases this may be due to the atomic radius of boron (87 pm) replaces the oxygen (48 pm atomic radius) in the TiO2 lattice. Boron present in TiO2 distorts the lattice and thereby increases the interplanar distance. This same phenomenon was observed in earlier reports [54,55].

2.5. X-ray Photoelectron Spectroscopy Studies

Figure 5 shows the X-ray photoelectron spectroscopy (XPS) spectra of 3B-TiO2-CNT nanocomposites. Figure 5a shows the survey spectrum which confirms the presence of Ti, O, C, and B elements in the synthesized nanocomposites. Figure 5b, Ti2p configuration, the spectra contain two symmetrical peaks centered at 458.20 eV and 464.01 eV, corresponding to Ti 2p3/2 and Ti 2p1/2, respectively, with a binding energy difference of 5.81 eV. It is recognized that the Ti element of nanocomposites exists in the Ti4+ oxidation state [34,56]. It is generally found that non-metal doping in TiO2 may be responsible for a change in the oxidation state from Ti4+ to Ti3+, but in the present research investigation, there is no Ti3+ peak, as shown in Figure 5b. The deconvoluted C 1s spectra are shown in Figure 5c. The three intense peaks at 284.64 eV, 286.34 eV, and 288.66 eV are attributed to graphitic carbon (C=C bonds), the C-C bonds of CNT, C-O bonds, and COO- bonds, respectively [57]. The O 1 s spectrum, shown in Figure 5d, contains two main peaks at 529.73 eV and 530.70 eV, which are respectively due to the oxygen in TiO2 and the surface hydroxyl groups present on the surface of TiO2 NPs [58]. Figure 5e shows the deconvoluted B 1s XPS spectrum of 3B-TiO2-CNT nanocomposites, which mainly includes peaks for B-O-B and B-Ti-O bonds, located at 192.14 eV and 189.66 eV, respectively [59,60,61]. The peak observed at 192.14 eV confirms that the boron dopant creates a band between the conduction and valence bands of pure TiO2, and the doped boron interacts with O2p orbitals of the TiO2 valence band. Therefore, Ti-O-B may have formed, which would confirm the doping of boron in the TiO2 host lattice [60].

2.6. Photodegradation of Toluene Gas

The photocatalytic degradation of toluene gas, under UV radiation in the presence of TiO2, 3B-TiO2 NPs, and B-doped TiO2-CNT nanocomposites, over a 120 min interval, is shown in Figure 6. Pure TiO2 NPs show less photocatalytic activity as compared to 3B-TiO2 and B-TiO2-CNT nanocomposites. The highest photocatalytic degradation efficiency, obtained for 3B-TiO2-CNT nanocomposites, was 96%, while pure TiO2 NPs, 3B-TiO2, 1B-TiO2-CNT, 2B-TiO2-CNT, and 4B-TiO2-CNT, showed photodegradation efficiencies of 37%, 51%, 76%, 83%, and 93% respectively. The impact of UV light on the photodegradation of toluene gas was studied in an experiment with no catalyst; UV light alone showed photodegradation efficiency of approximately 12%. In synthesized nanocomposites, optimized boron doping, as well as the addition of CNTs, playing a very important role in enhancing photocatalytic efficiency [62].
Pseudo-first-order reaction kinetics were observed for the photodegradation of toluene, as shown in Figure 7. 3B-TiO2-CNT nanocomposites showed the fastest photodegradation of toluene gas, having a rate constant k = 0.02528 min−1, while TiO2, 3B-TiO2 NPs, 1B-TiO2-CNT, 2B-TiO2-CNT, and 4B-TiO2-CNT had rate constants of k = 0.01950 min−1, k = 0.02250 min−1, k = 0.02130 min−1, k = 0.02170 and, k = 0.02370 respectively. The photodegradation reaction with UV light alone had a rate constant k = 0.01490 min−1.
In the photodegradation of toluene gas using B-TiO2-CNT nanocomposites, when UV light irradiated the sample, CNT and boron absorb light and electron excited from CNT goes to the conduction band (CB) of TiO2 while electron excited from valance band (VB) of TiO2 goes to the hole of CNT and boron band [63]. The hole formed in VB of TiO2 reacts with moisture to form hydroxyl radical and electron in CB of TiO2 reacts with oxygen to form superoxide. From the generated ROS, hydroxyl radical reacts with the phenyl ring of toluene to form various intermediates such as phenol, benzaldehyde, benzoic acid, and are finally converted into CO2 and H2O [64]. A possible photocatalytic degradation mechanism is shown in Figure 8.
To determine the stability of the photocatalyst, a recyclability test was performed with the same experimental setup used for the photodegradation of toluene. In this recyclability test, the 3B-TiO2-CNT nanocomposites were used as photocatalysts and irradiated with UV light for 120 min. After every cycle, the catalyst was removed from the Teflon bag and regenerated with heat treatment at 100 °C for 2 h. Then the regenerated catalyst was used again for the photodegradation of toluene. The catalyst was recycled four times, and the results are shown in Figure 9. Only a 4%–5% reduction in photodegradation efficiency was observed. This reduction in photodegradation efficiency may be due to the loss of catalyst in the regeneration process as well as in the transfer of catalyst powder from one place to another place. Overall, the experimental results show that the synthesized 3B-TiO2-CNT photocatalyst was highly stable and exhibited the best photocatalytic activity of all tested compounds up to four cycles.

2.7. Photocatalytic Inactivation of Bacteria

The photocatalytic inactivation activity of TiO2, 3B-TiO2, and B-TiO2-CNT nanocomposites was exhibited in the photocatalytic killing effect of E. coli, which was determined through a reduction in the number of colonies formed on Mueller–Hinton agar plates. As shown in Figure 10, the E. coli were almost entirely inactivated within 240 min by 3B-TiO2-CNT nanocomposites under visible radiation. Neither the photocatalyst by itself nor visible light alone affected the E. coli growth, therefore confirming that the observed inactivation is attributed to the photocatalytic reaction. The digital photographs of photoinactivation of E.coli with 3B-TiO2-CNT nanocomposite under visible light shown in Figure 11.
A possible mechanism of photocatalytic inactivation of E. coli is as follows when visible light irritated on B-TiO2-CNT nanocomposites. An electron from the CNT and boron band near to valance band of TiO2 get excited and transferred to the conduction band of TiO2. A hole formed on CNT attracts an electron from the valance band of TiO2 and the hole formed in TiO2 VB reacts with H2O to form hydroxyl radical. While electron present in CB of TiO2 reacts with oxygen to form superoxide. These formed ROS, especially hydroxyl radicals react with the E. coli cell wall and destruct the bacterial membrane and DNA strand, leading to cell lysis (release of metabolites, protein, and RNA from the cell and at the cell death occurs) [65,66].

3. Materials and Methods

3.1. Materials

Titanium (IV) butoxide and multi-wall carbon nanotubes (MWCNTs) were purchased from (Sigma-Aldrich, Saint Louis, MO, USA). Acetylacetone, boric acid, ethanol, extra pure sodium lauryl sulfate, were purchased from (Duksan, Gyeonggi-do, South Korea), and 10 ppm toluene gas purchased from (Korea Petrochemical Ind Co LTD, Seoul, South Korea).

3.2. Preparation of Boron-Doped TiO2-CNT Nanocomposites

A previously reported sol-gel method [49], with slight modifications, was used to prepare boron-doped TiO2-CNT nanocomposites with varying concentrations of boron from 1 mol% to 4 mol%. A 1:1 proportion of titanium (IV) butoxide to acetylacetone was added to a beaker. To this solution, an aqueous solution of sodium lauryl sulfate (capping agent; 2 wt.%) and excess distilled water were added with continuous stirring. Afterward, a stoichiometric concentration of boron (1 to 4 mol%), as an aqueous boric acid solution, was added, and the solution was vigorously stirred for 1 h at room temperature. Then 0.5 wt.% of CNT in aqueous solution were added. Afterward, the pH was adjusted to 8 with an ammonia solution, and the reaction mixture was stirred at 60 °C for 3 h, then allowed to cool to room temperature. The solution was filtered with Whatman filter paper, and the obtained precipitate was washed with ethanol and distilled water. The precipitate was dried in the oven and calcinated at 450 °C for 3 h. Pure TiO2 NPs and 3B-TiO2 were synthesized using the same method, omitting the boric acid and CNT respectively. The synthesized nanocomposites were denoted as TiO2, 3B-TiO2 NPs, 1B-TiO2-CNT, 2B-TiO2-CNT, 3B-TiO2-CNT, and 4B-TiO2-CNT, where the number before the B denotes the mol% of B present in the nanocomposites.

3.3. Photodegradation of Toluene Gas

Toluene gas was photodegraded by following a previously reported experimental method [67,68]. In this method, catalyst powder is uniformly spread in a polycarbonate petri dish and this petri dish is inserted into a 2 L Teflon film bag. Afterward, toluene gas with 10 ppm concentration was filled into the bag. adsorption-desorption equilibrium is created by keeping the experimental setup in the dark for 30 min. Then two 6 W UV light tubes, providing a total incident intensity of 2 mW/cm2 were used to irradiate the setup for a fixed time interval, after which 1 mL of toluene gas was withdrawn by syringe. The change in toluene gas concentration was examined using a gas chromatography (GC) system.

3.4. Photo-Inactivation of E. coli

The photocatalytic inhibition of E. coli was studied using a previously reported method [26], as follows. All glassware and samples were sterilized before use. E. coli (NCIM 2066) was cultured on a nutrient agar plate at 37 °C for 24 h. A 5 mL saline solution was used to prepare a bacterial suspension with a concentration of 106 CFU mL-1, and the photocatalyst concentration was 2 mg mL-1. Photocatalytic inhibition was carried out using TiO2, 3B-TiO2, and B-TiO2-CNT nanocomposites in the dark and under visible light. For visible light studies, a borosilicate glass reactor was illuminated with eight fluorescent tubes (Philips, 8 W, λ > 400 nm, light intensity ~0.5 mW cm−2) [48,69]. At specific time intervals, 100 μL aliquots of the irradiated bacterial suspension were withdrawn and spread on Mueller-Hinton agar plates. These plates were incubated at 37 °C for 24 h. The standard plate count method was used to determine viable numbers of cells in units of CFU mL−1 [69].

3.5. Characterization

The synthesized pure nanoparticles and nanocomposites were characterized using various techniques. A Bruker D8 Advance X-ray diffractometer (XRD) with a Cu K-alpha source (λ = 1.5406 Å) (Bruker, Aubrey, TX, USA) was used to study the phase and various structural parameters of the synthesized nanomaterials. A UV-3600 UV-vis spectrophotometer (Shimadzu, Kyoto, Japan) was used to measure the optical properties of the nanomaterials. The effect of the boron doping concentration on the recombination rate of electron-hole pairs was studied by photoluminescence (PL) with a Hitachi F-4500 fluorescence spectrophotometer (Hitachi, Tokyo, Japan). Transmission electron microscopy (TEM, Tecnai G2 Spirit TWIN, 120 kV, from FEI, Hillsboro, OR, USA) was used to study the morphological properties of the synthesized nanomaterials. The physical, chemical bonding, and elemental analyses of the nanocomposites were performed using X-ray photoelectron spectroscopy (XPS, Thermo Scientific K-Alpha, Waltham, MA, USA). The change in toluene gas concentration was determined by gas chromatography (GC, Agilent Technology 7890A, Santa Clara, CA, USA).

4. Conclusions

Boron-doped TiO2-CNT nanocomposites were synthesized successfully by the sol-gel method. It was observed that the fabrication of nanocomposites with CNT, coupled with boron doping in TiO2, positively affected the photocatalytic activity. XRD analysis confirmed that the anatase phase was present in the nanocomposites and yielded a calculated crystallite size of around 10–20 nm for B-TiO2-CNT nanocomposites. The uniform decoration of B-TiO2 nanoparticles on the CNT surface was confirmed by TEM. The slight increase in interplanar spacing confirms the presence of boron in the TiO2 host lattice. UV-DRS studies show red-shifted absorption bands for the nanocomposites. A qualitative XPS analysis recognized the Ti4+ oxidation state of the elemental Ti in the nanocomposites, as well as the boron doping, which created an intermediate band between the valence and conduction bands of TiO2. PL study shows that boron doping, as well as CNTs, plays an important role in reducing the recombination rate of electron-hole pair and enhancing photocatalytic efficiency. The photocatalytic activity was tested using toluene gas degradation and photoinactivation of E. coli; the results confirmed that the 3B-TiO2-CNT nanocomposites had superior photocatalytic efficiency to bare TiO2, 3B-TiO2 NPs, 1B-TiO2-CNT, 2B-TiO2-CNT, and 4B-TiO2-CNT. Recyclability studies confirmed that the 3B-TiO2-CNT nanocomposites were highly stable after four cycles. Therefore 3B-TiO2-CNT well suited for removing organic pollutants from the environment.

Author Contributions

J.-S.K. supervised the work. V.B.K. designed and performed the experiments. S.P.D. performed the antibacterial experiments. V.B.K., A.G.D., and S.-C.K. wrote the manuscript. All authors analyzed the results and reviewed the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Basic Study and Interdisciplinary R&D Foundation Fund of the University of Seoul 2019 (for Jung-Sik Kim).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nigar, H.; Julián, I.; Mallada, R.; Santamaría, J. Microwave-assisted catalytic combustion for the efficient continuous cleaning of VOC-containing air streams. Environ. Sci. Technol. 2018, 52, 5892–5901. [Google Scholar] [CrossRef] [PubMed]
  2. Park, J.-H.; Seo, Y.-S.; Kim, H.-S.; Kim, I.-K. Photodegradation of benzene, toluene, ethylbenzene, and xylene by fluidized bed gaseous reactor with TiO2/SiO2 photocatalysts. Korean J. Chem. Eng. 2011, 28, 1693–1697. [Google Scholar] [CrossRef]
  3. Sarigiannis, D.A.; Karakitsios, S.P.; Gotti, A.; Liakos, I.L.; Katsoyiannis, A. Exposure to major volatile organic compounds and carbonyls in European indoor environments and associated health risk. Environ. Int. 2011, 37, 743–765. [Google Scholar] [CrossRef]
  4. Pham, T.-D.; Lee, B.-K. Novel adsorption and photocatalytic oxidation for removal of gaseous toluene by V-doped TiO2/PU under visible light. J. Hazard. Mater. 2015, 300, 493–503. [Google Scholar] [CrossRef] [PubMed]
  5. Neubert, D.; Gericke, C.; Hanke, B.; Beckmann, G.; Baltes, M.M.; Kühl, K.-P.; Bochert, G.; Hartmann, J. Multicenter field trial on possible health effects of toluene. II. Cross-sectional evaluation of acute low-level exposure. Toxicology 2001, 168, 159–183. [Google Scholar] [CrossRef]
  6. Liou, T.-H.; Hung, L.-W.; Liu, C.-L.; Zhang, T.-Y. Direct synthesis of nano titania on highly-ordered mesoporous SBA-15 framework for enhancing adsorption and photocatalytic activity. J. Porous Mater. 2018, 25, 1337–1347. [Google Scholar] [CrossRef]
  7. Yang, C.-C.; Doong, R.-A.; Chen, K.-F.; Chen, G.-S.; Tsai, Y.-P. The photocatalytic degradation of methylene blue by green semiconductor films that is induced by irradiation by a light-emitting diode and visible light. J. Air Waste Manag. Assoc. 2018, 68, 29–38. [Google Scholar] [CrossRef] [Green Version]
  8. St. Denis, T.G.; Dai, T.; Izikson, L.; Astrakas, C.; Anderson, R.R.; Hamblin, M.R.; Tegos, G.P. All you need is light. Virulence 2011, 2, 509–520. [Google Scholar] [CrossRef]
  9. Byrne, A.J.; Dunlop, S.P.; Hamilton, W.J.; Fernández-Ibáñez, P.; Polo-López, I.; Sharma, K.P.; Vennard, S.A. A Review of Heterogeneous Photocatalysis for Water and Surface Disinfection. Molecules 2015, 20, 5574–5615. [Google Scholar] [CrossRef] [Green Version]
  10. McCullagh, C.; Robertson, J.M.C.; Bahnemann, D.W.; Robertson, P.K.J. The application of TiO2 photocatalysis for disinfection of water contaminated with pathogenic micro-organisms: A review. Res. Chem. Int. 2007, 33, 359–375. [Google Scholar] [CrossRef]
  11. Planas, O.; Bresolí-Obach, R.; Nos, J.; Gallavardin, T.; Ruiz-González, R.; Agut, M.; Nonell, S. Synthesis, photophysical characterization, and photoinduced antibacterial activity of methylene blue-loaded amino- and mannose-targeted mesoporous silica nanoparticles. Molecules 2015, 20, 6284–6298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Bartzatt, R.; Cirillo, S.; Cirillo, J. Antibacterial derivatives of ciprofloxacin to inhibit the growth of necrotizing fasciitis associated penicillin-resistant Escherichia coli. J. Pharm. 2013, 2013, 1–7. [Google Scholar]
  13. Bohara, R.A.; Pawar, S.H. Innovative developments in bacterial detection with magnetic nanoparticles. Appl. Biochem. Biotechnol. 2015, 176, 1044–1058. [Google Scholar] [CrossRef] [PubMed]
  14. Meshram, J.V.; Koli, V.B.; Phadatare, M.R.; Pawar, S.H. Anti-microbial surfaces: An approach for deposition of ZnO nanoparticles on PVA-Gelatin composite film by screen printing technique. Mater. Sci. Eng. C 2017, 73, 257–266. [Google Scholar] [CrossRef] [PubMed]
  15. Ziental, D.; Czarczynska-Goslinska, B.; Mlynarczyk, D.T.; Glowacka-Sobotta, A.; Stanisz, B.; Goslinski, T.; Sobotta, L. Titanium dioxide nanoparticles: Prospects and applications in medicine. Nanomaterials 2020, 10, 387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Fu, H.; Yang, L.; Hu, D.; Yu, C.; Ling, Y.; Xie, Y.; Li, S.; Zhao, J. Titanium dioxide nano-heterostructure with nanoparticles decorating nanowires for high-performance photocatalysis. Int. J. Hydrogen Energy 2018, 43, 10359–10367. [Google Scholar] [CrossRef]
  17. Bai, Y.; Mora-Seró, I.; De Angelis, F.; Bisquert, J.; Wang, P. Titanium dioxide nanomaterials for photovoltaic applications. Chem. Rev. 2014, 114, 10095–10130. [Google Scholar] [CrossRef]
  18. Marković, D.; Milovanovic, S.; Radovanović, Ž.; Zizovic, I.; Saponjic, Z.; Radetić, M. floating photocatalyst based on poly(ε-caprolactone) foam and TiO2 nanoparticles for removal of textile dyes. Fibers Polym. 2018, 19, 1219–1227. [Google Scholar] [CrossRef]
  19. Wang, Z.; Lang, X. Visible light photocatalysis of dye-sensitized TiO2: The selective aerobic oxidation of amines to imines. Appl. Catal. B Environ. 2018, 224, 404–409. [Google Scholar] [CrossRef]
  20. Eversfield, P.; Lange, T.; Hunger, M.; Klemm, E. Selective oxidation of o-xylene to phthalic anhydride on tungsten, tin, and potassium promoted VOx on TiO2 monolayer catalysts. Catal. Today 2019, 333, 120–126. [Google Scholar] [CrossRef]
  21. Ren, Y.; Sun, D.; Cao, Y.; Tsao, H.N.; Yuan, Y.; Zakeeruddin, S.M.; Wang, P.; Grätzel, M. A Stable blue photosensitizer for color palette of dye-sensitized solar cells reaching 12.6% efficiency. J. Am. Chem. Soc. 2018, 140, 2405–2408. [Google Scholar] [CrossRef] [PubMed]
  22. Lusvardi, G.; Barani, C.; Giubertoni, F.; Paganelli, G. Synthesis and characterization of TiO₂ nanoparticles for the reduction of water pollutants. Materials 2017, 10, 1208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Carlucci, C.; Xu, H.; Scremin, B.; Giannini, C.; Sibillano, T.; Carlino, E.; Videtta, V.; Gigli, G.; Ciccarella, G. Controllable one-pot synthesis of anatase TiO2 nanorods with the microwave-solvothermal method. Sci. Adv. Mater. 2014, 6, 1–8. [Google Scholar] [CrossRef]
  24. Yadav, H.M.; Otari, S.V.; Koli, V.B.; Mali, S.S.; Hong, C.K.; Pawar, S.H.; Delekar, S.D. Preparation and characterization of copper-doped anatase TiO2 nanoparticles with visible light photocatalytic antibacterial activity. J. Photochem. Photobiol. A Chem. 2014, 280, 32–38. [Google Scholar] [CrossRef]
  25. Barkul, R.P.; Koli, V.B.; Shewale, V.B.; Patil, M.K.; Delekar, S.D. Visible active nanocrystalline N-doped anatase TiO2 particles for photocatalytic mineralization studies. Mater. Chem. Phys. 2016, 173, 42–51. [Google Scholar] [CrossRef]
  26. Koli, V.B.; Delekar, S.D.; Pawar, S.H. Photoinactivation of bacteria by using Fe-doped TiO2-MWCNTs nanocomposites. J. Mater. Sci. Mater. Med. 2016, 27, 177. [Google Scholar] [CrossRef] [PubMed]
  27. Deshmukh, S.P.; Mullani, S.B.; Koli, V.B.; Patil, S.M.; Kasabe, P.J.; Dandge, P.B.; Pawar, S.A.; Delekar, S.D. Ag Nanoparticles Connected to the Surface of TiO2 Electrostatically for Antibacterial Photoinactivation Studies. Photochem. Photobiol. 2018, 94, 1249–1262. [Google Scholar] [CrossRef]
  28. Al Qarni, F.; Alomair, N.A.; Mohamed, H.H. Environment-friendly nanoporous titanium dioxide with enhanced photocatalytic activity. Catalysts 2019, 9, 799. [Google Scholar] [CrossRef] [Green Version]
  29. Olowoyo, J.O.; Kumar, M.; Jain, S.L.; Babalola, J.O.; Vorontsov, A.V.; Kumar, U. Insights into Reinforced Photocatalytic Activity of the CNT–TiO2 Nanocomposite for CO2 Reduction and Water Splitting. J. Phys. Chem. C 2019, 123, 367–378. [Google Scholar] [CrossRef]
  30. Gopannagari, M.; Kumar, D.P.; Park, H.; Kim, E.H.; Bhavani, P.; Reddy, D.A.; Kim, T.K. Influence of surface-functionalized multi-walled carbon nanotubes on CdS nanohybrids for effective photocatalytic hydrogen production. Appl. Catal. B Environ. 2018, 236, 294–303. [Google Scholar] [CrossRef]
  31. Blanchard, N.P.; Hatton, R.A.; Silva, S.R.P. Tuning the work function of surface oxidized multi-wall carbon nanotubes via cation exchange. Chem. Phys. Lett. 2007, 434, 92–95. [Google Scholar] [CrossRef]
  32. Wang, L.; Yao, Z.; Jia, F.; Chen, B.; Jiang, Z. A facile synthesis of ZnxCd1-xS/CNTs nanocomposite photocatalyst for H2 production. Dalton Trans. 2013, 42, 9976–9981. [Google Scholar] [CrossRef] [PubMed]
  33. Vijayan, B.K.; Dimitrijevic, N.M.; Finkelstein-Shapiro, D.; Wu, J.; Gray, K.A. Coupling titania nanotubes and carbon nanotubes to create photocatalytic nanocomposites. ACS Catal. 2012, 2, 223–229. [Google Scholar] [CrossRef]
  34. Quesada-González, M.; Boscher, N.D.; Carmalt, C.J.; Parkin, I.P. Interstitial boron-doped TiO2 thin films: The significant effect of boron on TiO2 coatings grown by atmospheric pressure chemical vapor deposition. ACS Appl. Mater. Interfaces 2016, 8, 25024–25029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Carlucci, C.; Scremin, B.F.; Sibillano, T.; Giannini, C.; Filippo, E.; Perulli, P.; Capodilupo, A.L.; Corrente, G.A.; Ciccarella, G. Microwave-assisted synthesis of boron-modified TiO2 nanocrystals. Inorganics 2014, 2, 264–277. [Google Scholar] [CrossRef] [Green Version]
  36. Hong, X.; Kang, Y.; Zhen, C.; Kang, X.D.; Yin, L.; Irvine, J.; Wang, L.; Liu, G.; Cheng, H.-M. Maximizing the visible light photoelectrochemical activity of B/N-doped anatase TiO2 microspheres with exposed dominant {001} facets. Sci. China Mater. 2018, 61, 831–838. [Google Scholar] [CrossRef] [Green Version]
  37. Sopchenski, L.; Cogo, S.; Dias-Ntipanyj, M.F.; Elifio-Espósito, S.; Popat, K.C.; Soares, P. Bioactive and antibacterial boron-doped TiO2 coating obtained by PEO. Appl. Surf. Sci. 2018, 458, 49–58. [Google Scholar] [CrossRef]
  38. Sohail, M.; Xue, H.; Jiao, Q.; Li, H.; Khan, K.; Wang, S.; Feng, C.; Zhao, Y. Synthesis of well-dispersed TiO2/CNTs@CoFe2O4 nanocomposites and their photocatalytic properties. Mater. Res. Bull. 2018, 101, 83–89. [Google Scholar] [CrossRef]
  39. Tan, T.L.; Lai, C.W.; Hong, S.L.; Rashid, S.A. New insights into the photocatalytic endocrine disruptors dimethyl phathalate esters degradation by UV/MWCNTs-TiO2 nanocomposites. J. Photochem. Photobiol. A Chem. 2018, 364, 177–189. [Google Scholar] [CrossRef]
  40. Akalework, N.; Pan, C.-J.; Su, W.-N.; Rick, J.; Tsai, M.-C.; Lee, J.-F.; Lin, J.-M.; Tsai, L.Y.; Hwang, B.J. Ultrathin TiO2-coated MWCNTs with excellent conductivity and SMSI nature as Pt catalyst support for oxygen reduction reaction in PEMFCs. J. Mater. Chem. 2012, 22, 20977–20985. [Google Scholar] [CrossRef]
  41. Koli, V.B.; Dhodamani, A.G.; Delekar, S.D.; Pawar, S.H. In situ sol-gel synthesis of anatase TiO2-MWCNTs nanocomposites and their photocatalytic applications. J. Photochem. Photobiol. A Chem. 2017, 333, 40–48. [Google Scholar] [CrossRef]
  42. Li, L.; Yang, Y.; Liu, X.; Fan, R.; Shi, Y.; Li, S.; Zhang, L.; Fan, X.; Tang, P.; Xu, R.; et al. A direct synthesis of B-doped TiO2 and its photocatalytic performance on the degradation of RhB. Appl. Surf. Sci. 2013, 265, 36–40. [Google Scholar] [CrossRef]
  43. Simsek, E.B. Solvothermal synthesized boron doped TiO2 catalysts: Photocatalytic degradation of endocrine disrupting compounds and pharmaceuticals under visible light irradiation. Appl. Catal. B Environ. 2017, 200, 309–322. [Google Scholar] [CrossRef]
  44. Szkoda, M.; Siuzdak, K.; Lisowska-Oleksiak, A. Non-metal doped TiO2 nanotube arrays for high efficiency photocatalytic decomposition of organic species in water. Phys. E Low-Dimens. Syst. Nanostruct. 2016, 84, 141–145. [Google Scholar] [CrossRef]
  45. Zhou, M.; Zhang, J.; Cheng, B.; Yu, H. Enhancement of visible-light photocatalytic activity of mesoporous Au-TiO2 nanocomposites by surface plasmon resonance. Int. J. Photoenergy 2012, 2012, 1–10. [Google Scholar]
  46. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-light photocatalysis in nitrogen-doped titanium oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef]
  47. Koli, V.; Dhodamani, A.; More, K.; Acquah, S.F.; Panda, D.K.; Pawar, S.; Delekar, S. A simple strategy for the anchoring of anatase titania on multi-walled carbon nanotubes for solar energy harvesting. Sol. Energy 2017, 149, 188–194. [Google Scholar] [CrossRef]
  48. Koli, V.B.; Dhodamani, A.G.; Raut, A.V.; Thorat, N.D.; Pawar, S.H.; Delekar, S.D. Visible light photo-induced antibacterial activity of TiO2-MWCNTs nanocomposites with varying the contents of MWCNTs. J. Photochem. Photobiol. A Chem. 2016, 328, 50–58. [Google Scholar] [CrossRef]
  49. Koli, V.B.; Mavengere, S.; Kim, J.-S. Boron-doped TiO2–CNTs nanocomposites for photocatalytic application. J. Mater. Sci. Mater. Electron. 2018, 29, 16660–16672. [Google Scholar] [CrossRef]
  50. Khan, R.; Kim, S.W.; Kim, T.-J.; Nam, C.-M. Comparative study of the photocatalytic performance of boron–iron Co-doped and boron-doped TiO2 nanoparticles. Mater. Chem. Phys. 2008, 112, 167–172. [Google Scholar] [CrossRef]
  51. Reddy, N.R.; Kumari, M.M.; Cheralathan, K.K.; Shankar, M.V. Enhanced photocatalytic hydrogen production activity of noble metal-free MWCNT-TiO2 nanocomposites. Int. J. Hydrogen Energy 2018, 43, 4036–4043. [Google Scholar] [CrossRef]
  52. Alwash, A.; Adil, H.; Hussain, Z.; Yousif, E. Potential of carbon nanotubes in enhance of photocatalyst activity. Arch. Nano Op. Acc. J. 2018, 1, 65–70. [Google Scholar]
  53. Schütt, F.; Signetti, S.; Krüger, H.; Röder, S.; Smazna, D.; Kaps, S.; Gorb, S.N.; Mishra, Y.K.; Pugno, N.M.; Adelung, R. Hierarchical self-entangled carbon nanotube tube networks. Nat. Commun. 2017, 8, 1–10. [Google Scholar] [CrossRef] [PubMed]
  54. Jeong, J.-H.; Jung, D.-w.; Shin, E.W.; Oh, E.-S. Boron-doped TiO2 anode materials for high-rate lithium ion batteries. J. Alloy. Compd. 2014, 604, 226–232. [Google Scholar] [CrossRef]
  55. Subramanian, A.; Wang, H.-W. Effects of boron doping in TiO2 nanotubes and the performance of dye-sensitized solar cells. Appl. Surf. Sci. 2012, 258, 6479–6484. [Google Scholar] [CrossRef]
  56. Zhang, W.; Yang, B.; Chen, J. Effects of calcination temperature on the preparation of boron-doped TiO2 by sol-gel method. Int. J. Photoenergy 2012, 2012, 1–8. [Google Scholar]
  57. Vilian, A.E.; Veeramani, V.; Chen, S.-M.; Madhu, R.; Kwak, C.H.; Huh, Y.S.; Han, Y.-K. Immobilization of myoglobin on Au nanoparticle-decorated carbon nanotube/polytyramine composite as a mediator-free H2O2 and nitrite biosensor. Sci. Rep. 2015, 5, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Yan, X.-b.; Tay, B.K.; Yang, Y. Dispersing and functionalizing multiwalled carbon nanotubes in TiO2 sol. J. Phys. Chem. B 2006, 110, 25844–25849. [Google Scholar] [CrossRef]
  59. Matos, J.; Laine, J.; Herrmann, J.-M.; Uzcategui, D.; Brito, J. Influence of activated carbon upon titania on aqueous photocatalytic consecutive runs of phenol photodegradation. Appl. Catal. B Environ. 2007, 70, 461–469. [Google Scholar] [CrossRef]
  60. Zhang, C.; Liu, Y.; Zhou, J.; Jin, W.; Chen, W. Tunability of photo-catalytic selectivity of B-doped anatase TiO2 microspheres in the visible light. Dyes Pigment. 2018, 156, 213–218. [Google Scholar] [CrossRef]
  61. Pan, H.; Zhang, Y.-W.; Shenoy, V.B.; Gao, H. Ab initio study on a novel photocatalyst: Functionalized graphitic carbon nitride nanotube. Acs Catal. 2011, 1, 99–104. [Google Scholar] [CrossRef]
  62. Fu, Y.; Chang, C.; Chen, P.; Chu, X.; Zhu, L. Enhanced photocatalytic performance of boron doped Bi2WO6 nanosheets under simulated solar light irradiation. J. Hazard. Mater. 2013, 254, 185–192. [Google Scholar] [CrossRef]
  63. Ghartavol, H.M.; Mohammadi, M.R.; Afshar, A.; Li, Y. On the assessment of incorporation of CNT–TiO2 core–shell structures into nanoparticle TiO2 photoanodes in dye-sensitized solar cells. Photochem. Photobiol. Sci. 2019, 18, 1840–1850. [Google Scholar] [CrossRef] [PubMed]
  64. Koli, V.B.; Mavengere, S.; Kim, J.-S. Photocatalytic properties of TiO2-SiO2-coated concrete on toluene gas. Mater. Res. Express 2018, 5, 125006. [Google Scholar] [CrossRef]
  65. Akhavan, O.; Abdolahad, M.; Abdi, Y.; Mohajerzadeh, S. Synthesis of titania/carbon nanotube heterojunction arrays for photoinactivation of E. coli in visible light irradiation. Carbon 2009, 47, 3280–3287. [Google Scholar] [CrossRef]
  66. Sunada, K.; Watanabe, T.; Hashimoto, K. Studies on photokilling of bacteria on TiO2 thin film. J. Photochem. Photobiol. A Chem. 2003, 156, 227–233. [Google Scholar] [CrossRef]
  67. Yadav, H.M.; Kim, J.-S. Solvothermal synthesis of anatase TiO2-graphene oxide nanocomposites and their photocatalytic performance. J. Alloy. Compd. 2016, 688, 123–129. [Google Scholar] [CrossRef]
  68. Koli, V.B.; Kim, J.-S. Photocatalytic oxidation for removal of gases toluene by TiO2-CeO2 nanocomposites under UV light irradiation. Mater. Sci. Semicond. Process. 2019, 94, 70–79. [Google Scholar] [CrossRef]
  69. Yadav, H.M.; Kolekar, T.V.; Pawar, S.H.; Kim, J.-S. Enhanced photocatalytic inactivation of bacteria on Fe-containing TiO2 nanoparticles under fluorescent light. J. Mater. Sci. Mater. Med. 2016, 27, 57. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of TiO2, 3B-TiO2, and B-TiO2-CNT nanocomposites.
Figure 1. XRD patterns of TiO2, 3B-TiO2, and B-TiO2-CNT nanocomposites.
Catalysts 10 00632 g001
Figure 2. (a) UV–vis DRS spectra of TiO2, 3B-TiO2, and B-TiO2-CNT nanocomposites; (b) Tauc plots for bandgap energy estimation of TiO2, 3B-TiO2, and B-TiO2-CNT nanocomposites.
Figure 2. (a) UV–vis DRS spectra of TiO2, 3B-TiO2, and B-TiO2-CNT nanocomposites; (b) Tauc plots for bandgap energy estimation of TiO2, 3B-TiO2, and B-TiO2-CNT nanocomposites.
Catalysts 10 00632 g002
Figure 3. PL spectra of TiO2, 3B-TiO2 and B-TiO2-CNT nanocomposites.
Figure 3. PL spectra of TiO2, 3B-TiO2 and B-TiO2-CNT nanocomposites.
Catalysts 10 00632 g003
Figure 4. TEM images of (a,b) TiO2, (c,d) 3B-TiO2, (e,f) 1B-TiO2-CNT and (g,h) 3B-TiO2-CNT nanocomposites.
Figure 4. TEM images of (a,b) TiO2, (c,d) 3B-TiO2, (e,f) 1B-TiO2-CNT and (g,h) 3B-TiO2-CNT nanocomposites.
Catalysts 10 00632 g004
Figure 5. XPS spectra of 3B-TiO2-CNT nanocomposites (a) survey spectrum (b) Ti 2p, (c) C 1s, (d) O 1s and (e) B 1s configuration.
Figure 5. XPS spectra of 3B-TiO2-CNT nanocomposites (a) survey spectrum (b) Ti 2p, (c) C 1s, (d) O 1s and (e) B 1s configuration.
Catalysts 10 00632 g005
Figure 6. Photocatalytic degradation of toluene gas under UV light irradiation.
Figure 6. Photocatalytic degradation of toluene gas under UV light irradiation.
Catalysts 10 00632 g006
Figure 7. The pseudo-first-order kinetics of degradation of toluene gas with a time of UV light irradiation.
Figure 7. The pseudo-first-order kinetics of degradation of toluene gas with a time of UV light irradiation.
Catalysts 10 00632 g007
Figure 8. Possible photocatalytic degradation mechanism of toluene gas with B-TiO2-CNT nanocomposites under UV light.
Figure 8. Possible photocatalytic degradation mechanism of toluene gas with B-TiO2-CNT nanocomposites under UV light.
Catalysts 10 00632 g008
Figure 9. Recyclability of photocatalyst 3B-TiO2-CNT nanocomposites.
Figure 9. Recyclability of photocatalyst 3B-TiO2-CNT nanocomposites.
Catalysts 10 00632 g009
Figure 10. % survival of E. coli with of TiO2, 3B-TiO2 and B-TiO2-CNT nanocomposites as a function of time.
Figure 10. % survival of E. coli with of TiO2, 3B-TiO2 and B-TiO2-CNT nanocomposites as a function of time.
Catalysts 10 00632 g010
Figure 11. Photographs of photo-inactivation of E. coli under visible light exposure.
Figure 11. Photographs of photo-inactivation of E. coli under visible light exposure.
Catalysts 10 00632 g011
Table 1. Crystallite size and optical bandgap of synthesized nanomaterials.
Table 1. Crystallite size and optical bandgap of synthesized nanomaterials.
Samples Crystallite Size (D) in nmOptical Band Gap (Eg) eV
TiO2103.2
3B-TiO2172.7
1B-TiO2-CNT123.0
2B-TiO2-CNT142.8
3B-TiO2-CNT172.6
4B-TiO2-CNT182.6

Share and Cite

MDPI and ACS Style

Koli, V.B.; Ke, S.-C.; Dodamani, A.G.; Deshmukh, S.P.; Kim, J.-S. Boron-Doped TiO2-CNT Nanocomposites with Improved Photocatalytic Efficiency toward Photodegradation of Toluene Gas and Photo-Inactivation of Escherichia coli. Catalysts 2020, 10, 632. https://doi.org/10.3390/catal10060632

AMA Style

Koli VB, Ke S-C, Dodamani AG, Deshmukh SP, Kim J-S. Boron-Doped TiO2-CNT Nanocomposites with Improved Photocatalytic Efficiency toward Photodegradation of Toluene Gas and Photo-Inactivation of Escherichia coli. Catalysts. 2020; 10(6):632. https://doi.org/10.3390/catal10060632

Chicago/Turabian Style

Koli, Valmiki B., Shyue-Chu Ke, Ananta G. Dodamani, Shamkumar P. Deshmukh, and Jung-Sik Kim. 2020. "Boron-Doped TiO2-CNT Nanocomposites with Improved Photocatalytic Efficiency toward Photodegradation of Toluene Gas and Photo-Inactivation of Escherichia coli" Catalysts 10, no. 6: 632. https://doi.org/10.3390/catal10060632

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop