Next Article in Journal
Iron Oxide Nanoparticles as Theranostic Agents in Cancer Immunotherapy
Previous Article in Journal
Clustering of Diamond Nanoparticles, Fluorination and Efficiency of Slow Neutron Reflectors
Previous Article in Special Issue
Iron Hydroxide/Oxide-Reduced Graphene Oxide Nanocomposite for Dual-Modality Photodynamic and Photothermal Therapy In Vitro and In Vivo
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Iron Modified Titanate Nanotube Arrays for Photoelectrochemical Removal of E. coli

1
Institute of Environment Engineering, National Sun Yat-sen University, Kaohsiung 804, Taiwan
2
Department of Marine Environmental Engineering, National Kaohsiung University of Science and Technology, Kaohsiung 811, Taiwan
3
Center for Environmental Toxin and Emerging-Contaminant Research, Cheng Shiu University, Kaoshiung 804, Taiwan
4
Department of Electrical Engineering, Cheng Shiu University, Kaoshiung 804, Taiwan
5
School of Environmental Science and Engineering, Guangdong University of Technology, Guangzhou 510006, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(8), 1944; https://doi.org/10.3390/nano11081944
Submission received: 29 June 2021 / Revised: 23 July 2021 / Accepted: 26 July 2021 / Published: 28 July 2021
(This article belongs to the Special Issue Nanomaterials for Photothermal Therapy and Antibacterial Applications)

Abstract

:
This study used iron modified titanate nanotube arrays (Fe/TNAs) to remove E. coli in a photoelectrochemical system. The Fe/TNAs was synthesized by the anodization method and followed by the square wave voltammetry electrochemical deposition (SWVE) method with ferric nitrate as the precursor. Fe/TNAs were characterized by SEM, XRD, XPS, and UV-vis DRS to investigate the surface properties and light absorption. As a result, the iron nanoparticles (NPs) were successfully deposited on the tubular structure of the TNAs, which showed the best light utilization. Moreover, the photoelectrochemical (PEC) properties of the Fe/TNAs were measured by current-light response and electrochemical impedance spectroscopy. The photocurrent of the Fe/TNAs-0.5 (3.5 mA/cm2) was higher than TNAs (2.0 mA/cm2) and electron lifetime of Fe/TNAs-0.5 (433.3 ms) were also longer than TNAs (290.3 ms). Compared to the photolytic (P), photocatalytic (PC), and electrochemical (EC) method, Fe/TNAs PEC showed the best removal efficiency for methyl orange degradation. Furthermore, the Fe/TNAs PEC system also performed better removal efficiency than that of photolysis method in E. coli degradation experiments.

1. Introduction

Titanium dioxide (TiO2) is one of the most promising photocatalysts because of its excellent properties such as resistance to acid and alkali corrosion, relative low price, non-toxicity, and small environmental footprint. Therefore, it has been widely applied in water treatment [1], atmospheric VOC (volatile organic compounds) removal [2], disinfection [3,4], hydrogen production [5,6,7], CO2 reduction [8], DSSC (dye-sensitized solar cell) [9,10], and sensing devices [11], etc. The photo-generated holes can react with water molecular to form hydroxyl radicals, which is a powerful oxidant, to oxidize most organic compounds such as dyes, PPCPs (pharmaceutical and personal care products) [6,12], VOCs [2], and PCDD/Fs (polychlorinated dibenzodioxins and polychlorinated dibenzofurans) [13,14]. On the other hand, the photo-generated electrons can react to generate hydrogen, which is a kind of clean energy. Moreover, the photo-generated electrons can also be applied for environmental applications, e.g., to reduce TCE (trichloroethylene) [15] and to react with ammonia borane (NH3BH3) to produce hydrogen [16].
The large band gap of 3.2 eV is the main obstacle for TiO2 applications. Many researchers employed non-metals such as N [17,18], C [19], F [20], and S [21] to dope into TiO2 to reduce the bandgap. The non-metal element can replace the oxygen vacancy to manipulate the lattice of TiO2 to form TiO2−xAx (where A refers to non-metal elements). The other way is applying metals and/or metal oxides such as Ag [6,22], Pt [23], Au [22,24], Pd [25], Fe [26], and Cu2O [5] to form a new Femi level of the as-synthesized metal-modified TiO2 to reduce the bandgap. Among various doping elements, trivalent iron is a transition metal that is easy to obtain, low in price, and difficult to react with organic compounds. The activity increase found for some reactions with Fe-doped titania has been attributed to a faster diffusion of reaction intermediates in comparison to pure TiO2 [27]. The bandgap of Fe-TiO2 was reduced from 3.08 to 2.2 eV after Fe modification and showed effective removal rate of phenol under 700 W visible light [28]. Zafar et al. (2021) used the iron doped titanium dioxide nanotubes (Fe-TNT) photocatalyst in various environmental water matrices including tap water, ultra-pure water, seawater, surface water, and deionized water under visible light. Results showed the Fe-TNT performed significant CR (Congo Red) decolorization, and TOC and COD removal in deionized water [29].
The other challenge of TiO2 application is the rapid recombination of electron-hole pairs. As the light with suitable energy (i.e., depends on the wavelength) TiO2, the electrons can overcome the bandgap to jump from the valence band to the conduction band, leading the holes in the valence band. The photo-generated electrons would recombine with electrons rapidly, therefore, reducing the oxidation and/or reduction ability of TiO2. By applying a bias potential between TiO2-anode and cathode in the photoelectrochemical system can significantly reduce the recombination of electron-hole pairs [30]. The electron lifetime can increase from 231.78 to 375.59 ms for silver modified TNAs than pure TNAs at +1.0 V (vs. Ag/AgCl) bias potential [6].
Bacteria is one of the indicators of water quality in wastewater treatment plants. To deal with bacteria issue, chlorination is the most used method for wastewater disinfection. However, chlorination disinfection methods cause concerns regarding the toxicity of chlorine residuals and its byproducts [31,32]. UV irradiation has become one of the most important alternatives to chlorination for wastewater disinfection nowadays. Tosa and Hirata (1999) found that the dose of UV light required for 90% and 99% inactivation of EHEC O157:H7 was 1.5 and 3.0 mW scm−2 [3]. Further UV disinfection combined with TiO2 [4], ozone [33], and radio frequency electric field [34] were also developed.
In this study, ferric nitrate was used as a precursor to modify titanium dioxide to improve the photocatalytic activity of titanium dioxide. The iron modified TNAs are characterized to demonstrate its physical, chemical and photoelectrochemical properties. The E. coli was elected as the bio-target for evaluating the PEC oxidation ability of Fe/TNAs for the very first time.

2. Materials and Methods

2.1. Preparation of Fe Modified Titanium Dioxide Nanotube Arrays (Fe/TNAs)

The titanium foil (99.5%, Zhang Jia Ltd., New Taipei City, Taiwan) was cut into the size of 2 × 2.5 cm and then was washed via acetone, ethanol, and DI water sequent to remove dust and impurities. After cleaning, the titanium foil was dried in the air. The pre-treated titanium foil was used to synthesize TNAs by anodization method. The positive electrode was connected to a titanium sheet, while the negative electrode was connected to a platinum (Pt) electrode in the solution of ammonium fluoride (0.5 wt % NH4F (98%, Alfa Aesar, Haverhill, MA, USA) and 6 vt % H2O) as the etching solution. The titanium foil was etched at a constant voltage of 40 V for 1 h to obtain TNAs. The as-synthesized TNAs was rinsed with DI water, and dried in oven at 40 °C. To increase the crystalline, the TNAs was calcined at 450 °C for 3 h. The iron deposited TNAs were synthesized by square wave voltammetry electrochemical deposition method (SWVE). First, pH of mixed solution of 0.2 M Fe(NO3)3·9H2O (98%, Alfa Aesar, Haverhill, MA, USA) and 0.05 M NaBH4 (99%, Alfa Aesar, Haverhill, MA, USA) was adjusted to 3.5 via 0.5 M HNO3 (65%, MERCK, Darmstadt, Germany). The TNAs sample was then immersed in the prepared solution for 20 min. The initial and final applied voltage was set as −1.0 and 0.0 V, respectively, with amplitude of 0.005 V and frequency of 2 Hz for the SWVE method. The as-synthesized samples were named as Fe/TNAs-0.2 and Fe/TNAs-0.5 according to the concentration of Fe(NO3)3·9H2O as 0.2 M and 0.5 M, respectively. The as-synthesized Fe/TNAs was rinsed by DI water and dried at 40 °C for two hours.

2.2. Characterization of Nanotube Arrays

The crystal structure of TNAs and Fe/TNAs was investigated by X-ray diffraction (XRD) (X’Pert Pro MRD, PANalytical, Almelo, The Netherlands) using a Cu Kα source at a wavelength of 0.154 nm. JCPDS PDF card database was selected as the identification of XRD peaks. The morphology was studied using a field-emission scanning electron microscope (Nova NanoSEM 430 FEI, Hillsboro, OR, USA). The specific surface area of TNAs and Fe/TNAs were measured by BET (ASAP 2020 N (S/N: 1195), Micromeritics, Norcross, GA, USA) analysis under 740 mmHg of pressure (P0) and 77.350 K of bath temperature. Raman analysis was conducted by using 532 nm Laser, with 1800 cm−1 grating and 50 s exposure time, operating by Labspec5 software. The X-ray photoelectron spectroscopy (XPS) experiments were conducted on the TNAs and Fe/TNAs using a PHI 5000 Versa Probe system (Physical Electronics, Chanhassen, MN, USA). The binding energy that was obtained from the XPS spectra was calibrated with reference to the C1s peak at 284.8 eV. The UV-vis absorption spectra were measured in diffused reflection mode using an integrating sphere (ISV-922, Jasco, Tokyo, Japan) that was attached to a Jasco V-750 UV-vis DRS spectrometer (V-750, Jasco, Tokyo, Japan).

2.3. PEC and Electrochemical Measurements

All PEC experiments were carried out in three-electrode mode at room temperature. A Fe/TNAs (2 cm2) and a Pt wire (99.997%, Alfa Aesar, Haverhill, MA, USA) served as a working electrode and a counter electrode, respectively. An Ag/AgCl (3 M KCl, ΩMetrohm, Herisau, Switzerland) electrode was selected as a reference electrode. A self-designed H-type reactor separated the anode and cathode to enhance the pollutants degradation. These two chambers were connected with a cation-exchanged membrane (Nafion 212, DuPount, Wilmington, DE, USA) to keep the ion balance in the system. A quartz window (7 cm2) on the side of the anode chamber provided high optical quality. For the photocurrent measurement, the operating light source is a 100 W mercury lamp (GGZ100, Shanghai Jiguang, Shanghai, China) at a bias potential of +1.0 V (vs. Ag/AgCl) in the electrolyte of 0.1 M NaCl (99%, Avantor, Radnor, PA, USA), while a platinum (Pt) and Ag/AgCl electrode work as the counter and reference electrode, respectively. Electrochemical impedance spectroscopy (EIS) was performed under an open circuit voltage with frequencies in the range from 10 kHz to 10 mHz with an AC voltage with an amplitude of 5 mV. The potential for I-V, I-t curve measurements and PEC degradation experiments were controlled by an electricity workstation (ΩMetrohm-Autolab B.V. PGSTAT204, Herisau, Switzerland). For the methyl orange (85%, Sigma Aldrich, St. Louis, Missouri, USA) degradation, the PEC, photocatalytic (PC), electrochemical (EC) and photolysis (P) experiments were examined at the illumination using 100 W Hg lamp. Applied voltage in PEC and EC process was +1.0 V (vs. Ag/AgCl). E. coli is a gram-negative bacteria and used as an indicator of fecal pollution in water. E. coli (CCRC 10674) was purchased from the Bioresource Collection and Research Center of the Food Industry Research and Development Institute (Hsinchu City, Taiwan). E. coli K-12 strain CCRC 10674 was grown aerobically in 100 mL of brain–heart infusion at 37 °C for 18 h and the micro-organisms were cultured at 25 °C. Consequently, 1 mL aqueous samples containing 70 CFU/mL initial E. coli concentration were prepared and subjected to treatment. In this study, E. coli was used to evaluate the Fe/TNAs PEC degradation ability. The operated light source is 100 W Hg lamp at a bias potential of +1.0 V (vs. Ag/AgCl), while a reference electrode of Ag/AgCl, counter electrode of Pt wire, and a working electrode of Fe/TNAs-0.5 were employed.

3. Results and Discussions

3.1. Surface Morphology of TNAs and Fe/TNAs

In this study, SEM was used to identify the morphology of the surface of the material, and to explore the tubular structure of TNAs and Fe/TNAs prepared by the anodic oxidation etching method. Figure 1 shows the SEM images of TNAs and Fe/TNAs, respectively. Figure 1a shows an array of titanium dioxide nanotubes synthesized by anodizing method. The structure is compact and neatly arranged, and the diameter of the tube were approximately in the range of 50–100 nm. Many previous studies have reported a linear correlation between the diameter of titanate nanotube arrays and the applied anode voltage (Marck et al., 2017; Yasuda et al., 2007). Figure 1b shows the cross-sectional view of the un-modified titanium dioxide nanotube arrays with a length about 1.3 μm. Cao et al. (2016) applied anodization method to synthesize titanium dioxide nanotube arrays and found that the length of TNAs increases with the time of anodization [35]. Figure 1c,d shows the iron modified titanate nanotube arrays synthesized (Fe/TNAs) via SWVE method with 0.2 and 0.5 M Fe(NO3)3·9H2O precursor, respectively. The structure of the nanotube is not damaged after ferric nitrate modification, and the iron is successfully deposited on the surface of the TNAs in the form of sol particles. The morphology of iron is irregular, and it is evenly distributed and accumulated around the surface of TNAs. Notably, as the concentration of precursor of iron nitrate increases, the iron deposits increased on the TNAs. Previous literature showed a similar deposition morphology, confirming that sodium borohydride can successfully reduce Fe3+ in ferric nitrate to zero-valent iron Fe0 [36]. Table 1 shows the results of EDX analyses of TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5, respectively. The measured elements include C, O, Ti, and Fe, individually. The atomic percentage of Fe were 0.00, 0.93, and 7.44% for TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5, respectively, illustrating that the Fe were successfully deposited on the surface of TNAs. Notably, as shown in Figure 1d, most surface of the TNAs were covered by Fe, therefore, the Ti content is much less in Fe-TNAs-0.5 (16.51%) than that of TNAs (34.15%) and Fe/TNAs-0.2 (33.84%). Figure S1 shows the BET nitrogen adsorption-desorption isotherms of TNAs and Fe/TNAs. The measured surface area of TNAs and Fe/TNAs were 0.1464 and 0.0029 m2/g, respectively.

3.2. XRD Analyses

Figure 2 shows the XRD results for TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5. Compared with the standard chart of the Joint Committee for Powder Diffraction Files (JCPDFs: 21–1272), the XRD results show that TNAs exhibits anatase characteristic peaks at 2θ = 25.3°, 37.8° and 54.0°, corresponding to the crystal planes of (101), (004), and (105), respectively. No rutile phase was observed due to the insufficient temperature [5,15]. Notably, Fe/TNAs-0.2 and Fe/TNAs-0.5 show the diffraction peaks of Fe at 2θ = 24.16°, 35.74°, 54.23°, 62.26°, corresponding to the crystal planes of (012), (110), (116), and (214) (JCPDS file number: 00-001-1053), respectively, indicating that the Fe nanoparticles have been deposited on the surface of TNAs. Similar results of XRD analyses for TiO2-Fe2O3 were found in literature, showing that the successfully deposited iron [37]. Moreover, with the increase amount of Fe (see Table 1), the diffraction peak intensities corresponding to Fe increases as well. Figure S2 shows the Raman analyses for bare TNAs and Fe/TNAs. In general, anatase phase of titanium dioxide has six Raman active modes 1A1g, 2B1g, and 3Eg. TNAs and Fe/TNAs show solo peak at 143.1 and 144.5 cm−1, respectively, that corresponding to the active mode of Eg of anatase phase of TiO2. The results in Ramana analyses are in good agreement with the XRD observations. Notably, the main Eg mode located at around 144 cm−1 broadened and shifted toward higher wavenumber in Fe/TNA (inset plot in Figure S1), indicating that the Fe3+ are successfully incorporated into the TiO2 framework, replacing Ti4+ cations. This results in in consistent to the findings of Fe-TiO2 NPs [38].

3.3. XPS Analyses

The electron energy spectrometer (XPS), as shown in Figure 3a–e, are used to clarify the composition of elements of Fe/TNAs and to analyze the energy spectrum. It is also used to identify the uniformity and dispersion of surface elements of materials. Figure 3a is the full-energy spectra of Fe/TNAs-0.5, showing that C, O, Ti, and Fe are distributed on the surface of the material. Figure 3b shows the characteristic peak of C 1s at 284.5 eV. The C element was attributed to the addition to the apparatus during the analysis. Figure 3c shows the characteristic peaks of Fe at 710 eV and 725 eV, corresponding to the element states of 2p3/2 and 2p1/2, respectively. Notably, the enhancement in binding energy in the elemental state of 2p3/2 is due to the fact that Fe3+ disperses to Ti4+ in the anatase lattice and forms Fe−O−Ti bonds [39]. Zue et al. (2006) used sol-gel method to synthesize Fe-TiO2 and indicated that the binding energies are 711.0–711.8 eV and 725.4–726.0 eV for 2p3/2 and 2p1/2, respectively, of Fe3+ [39], while Huerta-Flores, et al., (2021) illustrated that two main peaks were identified around 710.7 eV and 724.8 eV attributed to Fe 2p3/2 and 2p1/2 levels of Fe3+ in Fe2O3 material [40]. Above findings are different to the observations of Fe0 that showing the characteristic peaks at 707.8 and 720.2eV for Fe 2p3/2 and 2p1/2, respectively, in Wang et al. (2018) [41]. Hsieh et al. (2010) also indicated that two major peaks at 707.0 and 719.1 eV are representing the Fe 2p3/2 and 2p1/2, corresponding to the zero-valent iron [42]. Hence, we confirmed the state of Fe in Fe/TNAs is Fe3+ instead of Fe0 as the findings in Figure 3c. Figure 3d shows that O has a characteristic peak at 530 eV, which represents the formation of Ti or Fe oxides [43]. Figure 3e shows that the characteristic peaks of Ti at 458 eV and 464 eV correspond to Ti 2p3/2 and Ti 2p1/2 respectively, confirming that Ti exists in the form of Ti4+ [44]. Table 2 shows the element percentage of Fe/TNAs-0.2 and Fe/TNAs-0.5 by XPS analysis. The Fe content for Fe/TNAs-0.2 and Fe/TNAs-0.5 is 0.9% and 2.9%, respectively, that are less than the results of EDX analyses. This is because the distribution of iron particles on the surface of the titanium sheet is uneven, and the particle size is different. The comparison of element components showed a slight difference between EDX and XPS because of the following reasons: (a) the range of analytical depth for XPS and EDX is within 10 and 100–1000 nm, respectively; (b) the analytical area of samples was larger in XPS than that in EDX; (c) the appearance of the C element can be attributed to extrinsic carbon during the sample fabrication and/or the XPS instrument itself.

3.4. UV-Vis DRS Analyses

The absorption coefficient of the thin films was calculated with the following formula:
α ( h v ) = 1 d ln ( T ( 1 R ) 2 )
where α is the absorption coefficient, d is the thickness of the film, and for a particular wavelength T is transmission and R is the reflectance characteristics of the film.
A ( h v E G ) = ( α h v ) 1 n
The optical band gap is calculated by plotting (αhv)1/n versus hv (i.e., eV), where α is the absorption coefficient of the material, h is the Planck constant, ν is the photon’s frequency, A is a constant, and EG is the bandgap energy. As to the constant n is 0.5 and 1 for indirect and direct band gap materials, respectively. TiO2 has an indirect band gap [38]. As shown in Figure 4, the absorbance wavelength is 420, 530, 730 nm for TNAs, Fe/TNAs-0.2 and Fe/TNAs-0.5, respectively. The absorbance wavelength of TNAs red shift to longer wavelength region after Fe nanoparticles modification illustrating the better light utilization. Band gap (EG) can be calculated according to the Tauc formula [45] as shown Equation (2) above. The inset plot in Figure 4 shows the corresponding bandgap is 2.9, 2.3, and 1.7 eV for TNAs, Fe/TNAs-0.2 and Fe/TNAs-0.5, individually. The results here are similar to the findings of Fe-TiO2 thin film [46] and Fe-TiO2 nanoparticles [38,47] that applying Tauc formula to investigate the bandgap energy. In general, there is a good correlation between light absorption and photocatalytic activity. Stronger light absorption correlated with higher photocatalytic activity. UV-vis DRS spectra results confirmed that Fe nanoparticles were successfully deposited on TNAs.

3.5. Photocurrent Measurement

Figure 5 shows the measured photocurrent diagram by using TNAs, Fe/TNAs-0.2 and Fe/TNAs-0.5 as the photo-anode in a PEC system. As shown in the figure, the photocurrent was monitored under a 50 s on-off cycle. The photocurrent of three tested photo-anodes is about 1–4 μA when the illumination is off. Because there is no light to excite the photoelectrons on the surface of the material, leading to no photocurrent is generated. When the light is on, all materials generate photocurrent as a stable straight line, which means that the photoelectrochemical performance of the photo-anodes tend to a stable state. The photocurrent densities of TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5 are 2.0, 3.0, and 3.5 mA/cm2, respectively. Notably, the photocurrent was 1.75 times higher for Fe/TNAs-0.5 than that of pure TNAs at the same illumination. TNAs modified by iron can excite more photo-generated electrons than TNAs under irradiation, which promotes the increase of photocurrent, indicating that more electron-hole pairs are generated on the surface of the material, resulting to better degrade pollutants.

3.6. Electrochemical Impedance Spectrum

In this study, electrochemical impedance spectrum (EIS) was used to analyze the electron-hole separation ability and electron transfer characteristics of TNAs, Fe/TNAs-0.2 and Fe/TNAs-0.5 under light irradiation. The smaller the diameter of the arc refers to the better the separation of electrons and holes, resulting in better photoelectrochemical ability. Figure 6a shows the Nyquist plots of TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5 by EIS analysis, indicating that the diameter of each semicircle follows the order: TNAs > Fe/TNAs-0.2 > Fe/TNAs-0.5, which demonstrating that the electron mobility of TNAs was enhanced by Fe modification. Figure 6b shows the Bode plot obtained by EIS analysis for three photo-anodes. The maximum frequency obtained from the figure is then fitted with the equation of τ e l = 1 2 π × f m a x to calculate the electron lifetime. Table 3 shows the electron lifetime was 290.3, 354.7, and 433.3 ms, corresponding to the fmax of 0.548, 0.449, and 0.367 Hz in Figure 6a, for TNA, Fe/TNAs-0.2, and Fe/TNAs-0.5, respectively. Table 3 summarizes the result of fitted equivalent circuit for TNA, Fe/TNAs-0.2, and Fe/TNAs-0.5, respectively. Notably, Rs refers to the resistance of the solution and Rp represents the charge transfer at the interface between Fe/TNAs and electrolyte, while the CPE refers to the constant phase element [5]. The Rp was 290.5, 143.33, and 130.97 for TNA, Fe/TNAs-0.2, and Fe/TNAs-0.5, individually, with +1.0 V (vs. Ag/AgCl) of bias potential. According to the results of Figure 6 and Table 3, we can preliminarily conclude that TNAs modified with iron can effectively reduce the recombination of electron holes, prolong the residence time of electrons or increase the mobility of electrons in the PEC system.

3.7. Mott-Schottky Plot for TNAs and Fe/TNAs

The Mott-Schottky plot is performed by applying the electrochemical impedance method, which is carried out in the range of −0.6 V to +0.2 V (vs. Ag/AgCl) under frequency of 100 Hz and amplitude of 10 mV. The type and flat band potential of semiconductor were determined by following equation [40,48,49]:
1 C 2 = 2 ε ε 0 A 2 q N D ( V V f b K T q )
where ε is the dielectric constant of TiO2 (i.e., 31 for anatase) [50,51], ε0 is the permittivity of vacuum (8.85 × 10−14 F cm−1), A is the area of photo-anode (4 cm2), q is the elementary charge (1.6 × 10−19 C), ND is the density of dopants (cm−3), V is the applied voltage, Vfb is the flat band potential, K is the Boltzman constant (1.381 × 10−23 J K−1), and T is the absolute temperature (298 K). In addition, the charge carrier density (ND) of semiconductor was also calculated from the slope of Mott–Schottky plots, which is following equation [48,52,53]:
d C 2 d V = 2 q ε ε 0 N D A 2
As shown in Figure 7, both TNAs and Fe/TNAs show the positive slope, confirming that Fe depositing on TNAs does not change its n-type semiconductor property. Moreover, the Vfb of TNAs and Fe/TNAs was −0.34 and −0.25 V (vs. Ag/AgCl), individually, which were estimated from the intercept with the X-axis on the linear plot. The positive shift of Vfb in Fe/TNAs photo-anode implies the facilitating of photo-generated electron-hole pairs separation and transfer, which is corresponding to EIS results (Figure 6) [48]. The charge carrier density (ND) of TNAs (9.04 × 1021) was higher than that of Fe/TNAs (2.52 × 1021 cm−3). This was attributed to the slope of Fe/TNAs photo-anode (1.13 × 108) which was higher than that of TNAs photo-anode (3.15 × 107), indicating the significant decrease of charge carrier density of Fe/TNAs. Similar observations were found in Freitas et al. (2014), indicating that the increasing of charge carrier density was attributed to a high level of defects caused by oxygen vacancies [54].
In addition, when photo-anode is contacted with the electrolyte, the electrons on photo-anode will transfer to electrolyte spontaneously to form depletion region with positive charge between the interface of photo-anode and electrolyte. The width of depletion region (W) was also derived from the Mott–Schottky plot relationship and is described by following equation [48,54]:
W = ( 2 ε ε 0 ( V V f b ) q N D ) 1 2
The depletion region of TNAs and Fe/TNAs was 0.311 and 0.661 nm, individually. As the result, the oxygen vacancies were decreased with the substitution of Ti4+ by Fe3+ at the surface, enlarging the space charge region at TiO2-electrolyte interface [48]. Furthermore, the conduction band (ECB) is determined from the following equation for n-type semiconductor:
E C B = V f b + K T   ln ( N D N c )
where K is the Bolzman constant (8.617 × 10−5 eV K−1) and Nc is the effective density of states (DOS) at the conduction band edge, which can be calculated by using equation: N c = 2 ( 2 π   m e f f K T h 2 ) 3 2 . The meff of 9 m0 (for TiO2 anatase) is used for Nc calculations [55], where m0 is the mass of electron (9.109 × 10−31 kg), K is the Bolzman constant (1.381 × 10−23 J K−1), T is the absolute temperature (298 K), and h is the Planck constant (6.626 × 10−34 J s). By applying the equation, Nc is determined as 6.71 × 1020 cm−3. Therefore, the conduction band −0.27 and −0.22 V for TNAs and Fe/TNAs, respectively. Furthermore, UV-vis DRS results showed that the band gap of TNAs and Fe/TNAs was 2.9 and 1.7 eV, respectively, thus, the position of the valence band was at approximately 2.63 and 1.48 V for TNAs and Fe/TNAs, respectively.

3.8. Methyl Orange Removal

Four methyl orange removal processes, namely, PEC, PC, EC and P processes were conducted to evaluate the degradation efficiency of TNA, Fe/TNAs-0.2, and Fe/TNAs-0.5. As shown in Figure 8, PEC process was the most efficient way to degrade MO among three methods studied. The complete removal of 10 ppm of MO was observed after 120, 50 and 30 min for the PEC method by using TNA, Fe/TNAs-0.2, and Fe/TNAs-0.5, respectively. The order of MO degradation by TNA, Fe/TNAs-0.2, and Fe/TNAs-0.5 is in consistent to the photocurrent densities of TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5 that are 2.0, 3.0, and 3.5 mA/cm2, respectively. According to the EIS analyses, the electron lifetime was 290.3, 354.7, and 433.3 ms for TNA, Fe/TNAs-0.2, and Fe/TNAs-0.5, respectively, illustrating that the freedom of electrons has increased. This enhanced PEC performance was attributed to the Fe modified as evidence of results of EDX and XPS.
Briefly, the MO degradation efficiency in four methods follows: PEC > PC > P > EC. It was observed that the PEC method provided the most powerful way to degrade MO due to the combination of electrochemical oxidation and photocatalysis [5].

3.9. E. coli Removal by Fe/TNAs PEC System

E. coli was selected as the bio-target for investigating the PEC disinfection ability. Figure 8 shows the E. coli removal efficiency was 87.02% by using Fe/TNAs-0.5 as the photo-anode in the PEC system within the reaction time of 60 min. E. coli degradation by photolysis, the same light source (100 W Hg lamp) in the PEC, was conducted for comparison. The images of E. coli removal by TNAs and Fe/TNAs PEC system were demonstrated in Figure S3. As shown in Figure 9, only 33.1% of E. coli was removed in the photolysis procedure. It is well known that UV light is commonly used in traditional water treatment system. According to the literature, inactivation of EHEC O26 with photoreactivation is 2.2 times greater than that without photoreactivation [3]. In addition, with the addition of TiO2 in UV disinfection system, the damage of outer membrane was observed for the cells. TiO2 alone can break down lipopolysaccharide (LPS), the outermost layer of the E. coli cells [4]. The hydroxyl radicals, which is a kind of most strong oxidants, were responsible for E. coli removal in the TNAs PEC system [6]. Compared to photolysis at the same irradiation, PEC system can generate strong oxidants, i.e., hydroxyl radicals [5,6], to degrade E. coli. On the other hand, the recombination of electron-hole pairs is greatly reduced in PEC method as evidence of enhanced electron lifetime (see Table 3), that allowed more photo-generated holes to react with H2O to form OH⋅ to removal E. coli. The result is in consistent to the findings of Ma et al. (2021) that Z-scheme g-C3N4/TNAs reduced recombination of photo-generated electron-hole pairs [56].

4. Conclusions

Fe/TNAs were fabricated by the SWVE method to remove the E. coli for the very first time in this study. The SEM results showed that iron NPs were deposited on the surface of the TNAs in the form of sol particles and the tubular structure was not damaged after iron modification. The Fe content on the surface of TNAs was increased with the increasing of the precursor concentration, which demonstrated by EDX and XPS results. XPS also revealed that the binding energy of Fe located at 710 and 725 eV, which represent +3 state of Fe. XRD analyses showed the characteristic peak of Fe at 2θ = 24.16°, 35.74°, 54.23°, and 62.26°, corresponding to the crystal planes of (012), (110), (116), and (214), illustrating the successful loading of iron NPs. Moreover, UV-vis DRS showed that the light absorbance was increased and the bandgap energy was decreased after iron modification to enhance the light utilization. The results of photoelectrochemical performance indicated that Fe/TNAs-0.5 performed the best photocurrent density (3.5 mA/cm2) as well as the electron lifetime of Fe/TNAs-0.5 is 1.49 times greater than that of pure TNAs through the EIS and Bode plot analyses, illustrating that the loading of Fe NPs is advantage to the separation of electron-hole pairs to enhance the photoelectrochemical performance. The deep and shallow trap states in bandgap of TNAs play an essential role in the electron mobility and PEC performances. Therefore, investigation of the effect of the traps of the TNAs is recommended for future study. Compared to the P, PC, and EC methods, PEC was the best method for MO degradation by using Fe/TNAs-0.5 as the photo-anode. Furthermore, the Fe/TNAs-0.5 PEC system also performed better removal efficiency than the P method for E. coli degradation. To sum up, the iron NPs deposited TNAs were successfully synthesized in this study. Depositing Fe NPs on TNAs not only reduces the bandgap energy of the TNAs but also decreases the electron-hole recombination, therefore, leading to the enhancement of the PEC performances in MO and E. coli removal.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano11081944/s1, Figure S1: BET analysis of (a) TNAs and (b) Fe/TNAs, Figure S2: Raman spectrum of TNAs and Fe/TNAs, Figure S3: Images of E. coli removal in photolysis and PEC system.

Author Contributions

Conceptualization, Y.-P.P. and Y.-C.L.; methodology, Y.-P.P. and C.-H.C.; formal analysis, M.-H.L. and C.-H.C.; investigation, K.-L.C. and C.-H.C.; writing—original draft preparation, Y.-P.P.; writing—review and editing, Y.-P.P., Y.-C.L., J.S. and C.-H.C.; supervision, Y.-P.P. and Y.-C.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology, Taiwan, under Grant No. MOST 108-2628-E-110-007-MY3 and the Environmental Protection Administration (EPA), Taiwan.

Institutional Review Board Statement

This study did not involve humans or animals.

Informed Consent Statement

This study did not involve humans.

Data Availability Statement

Data is contained within article.

Acknowledgments

This study was funded by the Ministry of Science and Technology, Taiwan, under Grant No. MOST 108-2628-E-110-007-MY3 and the Environmental Protection Administration (EPA), Taiwan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lee, S.-Y.; Park, S.-J. TiO2 photocatalyst for water treatment applications. J. Ind. Eng. Chem. 2013, 19, 1761–1769. [Google Scholar] [CrossRef]
  2. Zou, L.; Luo, Y.; Hooper, M.; Hu, E. Removal of VOCs by photocatalysis process using adsorption enhanced TiO2–SiO2 catalyst. Chem. Eng. Process. Process Intensif. 2006, 45, 959–964. [Google Scholar] [CrossRef]
  3. Tosa, K.; Hirata, T. Photoreactivation of enterohemorrhagic Escherichia coli following UV disinfection. Water Res. 1999, 33, 361–366. [Google Scholar] [CrossRef]
  4. Liu, P.; Duan, W.; Wang, Q.; Li, X. The damage of outer membrane of Escherichia coli in the presence of TiO2 combined with UV light. Coll. Surf. B Biointerfaces 2010, 78, 171–176. [Google Scholar] [CrossRef]
  5. Sun, Q.; Peng, Y.-P.; Chen, H.; Chang, K.-L.; Qiu, Y.-N.; Lai, S.-W. Photoelectrochemical oxidation of ibuprofen via Cu2O-doped TiO2 nanotube arrays. J. Hazard. Mater. 2016, 319, 121–129. [Google Scholar] [CrossRef]
  6. Peng, Y.-P.; Liu, C.-C.; Chen, K.-F.; Huang, C.-P.; Chen, C.-H. Green synthesis of nano-silver–titanium nanotube array (Ag/TNA) composite for concurrent ibuprofen degradation and hydrogen generation. Chemosphere 2021, 264, 128407. [Google Scholar] [CrossRef]
  7. Üzer, E.; Kumar, P.; Kisslinger, R.; Kar, P.; Thakur, U.K.; Zeng, S.; Shankar, K.; Nilges, T.J. Vapor deposition of semiconducting phosphorus allotropes into TiO2 nanotube arrays for photoelectrocatalytic water splitting. ACS Appl. Nano Mater. 2019, 2, 3358–3367. [Google Scholar] [CrossRef]
  8. Zeng, S.; Vahidzadeh, E.; VanEssen, C.G.; Kar, P.; Kisslinger, R.; Goswami, A.; Zhang, Y.; Mahdi, N.; Riddell, S.; Kobryn, A.E. Optical control of selectivity of high rate CO2 photoreduction via interband-or hot electron Z-scheme reaction pathways in Au-TiO2 plasmonic photonic crystal photocatalyst. Appl. Catal. B Environ. 2020, 267, 118644. [Google Scholar] [CrossRef]
  9. Lee, K.-H.; Han, S.-H.; Chuquer, A.; Yang, H.-Y.; Kim, J.; Pham, X.-H.; Yun, W.-J.; Jun, B.-H.; Rho, W.-Y. Effect of Au Nanoparticles and Scattering Layer in Dye-Sensitized Solar Cells Based on Freestanding TiO2 Nanotube Arrays. Nanomaterials 2021, 11, 328. [Google Scholar] [CrossRef]
  10. Mor, G.K.; Shankar, K.; Paulose, M.; Varghese, O.K.; Grimes, C.A. Use of highly-ordered TiO2 nanotube arrays in dye-sensitized solar cells. Nano Lett. 2006, 6, 215–218. [Google Scholar] [CrossRef] [PubMed]
  11. Zarifi, M.; Mohammadpour, A.; Farsinezhad, S.; Wiltshire, B.; Nosrati, M.; Askar, A.; Daneshmand, M.; Shankar, K. Time-resolved microwave photoconductivity (TRMC) using planar microwave resonators: Application to the study of long-lived charge pairs in photoexcited titania nanotube arrays. J. Phys. Chem. C 2015, 119, 14358–14365. [Google Scholar] [CrossRef]
  12. Chen, H.; Peng, Y.-P.; Chen, T.-Y.; Chen, K.-F.; Chang, K.-L.; Dang, Z.; Lu, G.-N.; He, H. Enhanced photoelectrochemical degradation of Ibuprofen and generation of hydrogen via BiOI-deposited TiO2 nanotube arrays. Sci. Total Environ. 2018, 633, 1198–1205. [Google Scholar] [CrossRef] [PubMed]
  13. Weber, R.; Sakurai, T.; Hagenmaier, H. Low temperature decomposition of PCDD/PCDF, chlorobenzenes and PAHs by TiO2-based V2O5–WO3 catalysts. Appl. Catal. B Environ. 1999, 20, 249–256. [Google Scholar] [CrossRef]
  14. Gallastegi-Villa, M.; Aranzabal, A.; Gonzalez-Marcos, J.; Gonzalez-Velasco, J. Tailoring dual redox-acid functionalities in VOx/TiO2/ZSM5 catalyst for simultaneous abatement of PCDD/Fs and NOx from municipal solid waste incineration. Appl. Catal. B Environ. 2017, 205, 310–318. [Google Scholar] [CrossRef]
  15. Peng, Y.-P.; Peng, L.-C.; Chen, K.-F.; Chen, C.-H.; Chang, K.-L.; Chen, K.-S.; Dang, Z.; Lu, G.-N.; Sun, J. Degradation of trichloroethylene by photoelectrochemically activated persulfate. Chemosphere 2020, 254, 126796. [Google Scholar] [CrossRef]
  16. Komova, O.; Simagina, V.; Kayl, N.; Odegova, G.; Netskina, O.; Chesalov, Y.; Ozerova, A. Improved low-temperature hydrogen generation from NH3BH3 and TiO2 composites pretreated with water. Int. J. Hydrogen Energy 2013, 38, 6442–6449. [Google Scholar] [CrossRef]
  17. Peng, Y.-P.; Lo, S.-L.; Ou, H.-H.; Lai, S.-W. Microwave-assisted hydrothermal synthesis of N-doped titanate nanotubes for visible-light-responsive photocatalysis. J. Hazard. Mater. 2010, 183, 754–758. [Google Scholar] [CrossRef] [PubMed]
  18. Vaiano, V.; Sacco, O.; Sannino, D.; Ciambelli, P. Photocatalytic removal of spiramycin from wastewater under visible light with N-doped TiO2 photocatalysts. Chem. Eng. J. 2015, 261, 3–8. [Google Scholar] [CrossRef]
  19. Wang, H.; Lewis, J.P. Effects of dopant states on photoactivity in carbon-doped TiO2. J. Phys. Condens. Matter 2005, 17, L209. [Google Scholar] [CrossRef]
  20. Samsudin, E.M.; Abd Hamid, S.B.; Juan, J.C.; Basirun, W.J.; Centi, G. Synergetic effects in novel hydrogenated F-doped TiO2 photocatalysts. Appl. Surf. Sci. 2016, 370, 380–393. [Google Scholar] [CrossRef]
  21. Ohno, T.; Akiyoshi, M.; Umebayashi, T.; Asai, K.; Mitsui, T.; Matsumura, M. Preparation of S-doped TiO2 photocatalysts and their photocatalytic activities under visible light. Appl. Catal. A Gen. 2004, 265, 115–121. [Google Scholar] [CrossRef]
  22. Chan, S.C.; Barteau, M.A. Preparation of highly uniform Ag/TiO2 and Au/TiO2 supported nanoparticle catalysts by photodeposition. Langmuir 2005, 21, 5588–5595. [Google Scholar] [CrossRef]
  23. Kar, P.; Zhang, Y.; Mahdi, N.; Thakur, U.K.; Wiltshire, B.D.; Kisslinger, R.; Shankar, K. Heterojunctions of mixed phase TiO2 nanotubes with Cu, CuPt, and Pt nanoparticles: Interfacial band alignment and visible light photoelectrochemical activity. Nanotechnology 2017, 29, 014002. [Google Scholar] [CrossRef]
  24. Li, H.; Bian, Z.; Zhu, J.; Huo, Y.; Li, H.; Lu, Y. Mesoporous Au/TiO2 nanocomposites with enhanced photocatalytic activity. J. Am. Chem. Soc. 2007, 129, 4538–4539. [Google Scholar] [CrossRef]
  25. Lin, Y.-C.; Chen, C.-H.; Chen, K.-S.; Peng, Y.-P.; Lin, Y.-C.; Huang, S.-W.; Huang, C.-E.; Lai, H.-W.; Li, H.-W. Green Synthesized Palladium Coated Titanium Nanotube Arrays for Simultaneous Azo-Dye Degradation and Hydrogen Production. Catalysts 2020, 10, 1330. [Google Scholar] [CrossRef]
  26. Dolat, D.; Mozia, S.; Ohtani, B.; Morawski, A. Nitrogen, iron-single modified (N-TiO2, Fe-TiO2) and co-modified (Fe, N-TiO2) rutile titanium dioxide as visible-light active photocatalysts. Chem. Eng. J. 2013, 225, 358–364. [Google Scholar] [CrossRef]
  27. Arana, J.; Dıaz, O.G.; Saracho, M.M.; Rodrıguez, J.D.; Melián, J.H.; Peña, J.P. Photocatalytic degradation of formic acid using Fe/TiO2 catalysts: The role of Fe3+/Fe2+ ions in the degradation mechanism. Appl. Catal. B Environ. 2001, 32, 49–61. [Google Scholar] [CrossRef]
  28. Liu, Y.; Zhou, S.; Yang, F.; Qin, H.; Kong, Y. Degradation of phenol in industrial wastewater over the F–Fe/TiO2 photocatalysts under visible light illumination. Chin. J. Chem. Eng. 2016, 24, 1712–1718. [Google Scholar] [CrossRef]
  29. Zafar, Z.; Fatima, R.; Kim, J.-O.J.E.R. Experimental studies on water matrix and influence of textile effluents on photocatalytic degradation of organic wastewater using Fe–TiO2 nanotubes: Towards commercial application. Environ. Res. 2021, 197, 111120. [Google Scholar] [CrossRef] [PubMed]
  30. Mor, G.K.; Varghese, O.K.; Paulose, M.; Shankar, K.; Grimes, C.A. A review on highly ordered, vertically oriented TiO2 nanotube arrays: Fabrication, material properties, and solar energy applications. Sol. Energy Mater. Sol. Cells 2006, 90, 2011–2075. [Google Scholar] [CrossRef]
  31. Federation, W.E. Water Environment Federation Ultraviolet Disinfection; Federation, W.E., Ed.; Water Environment Federation: Alexandria, VA, USA, 1996; pp. 227–291. [Google Scholar]
  32. Chen, W.-H.; Wang, Y.-H.; Hsu, T.-H. The competitive effect of different chlorination disinfection methods and additional inorganic nitrogen on nitrosamine formation from aromatic and heterocyclic amine-containing pharmaceuticals. Chemosphere 2021, 267, 128922. [Google Scholar] [CrossRef] [PubMed]
  33. Bialka, K.L.; Demirci, A.J. Decontamination of Escherichia coli O157: H7 and Salmonella enterica on blueberries using ozone and pulsed UV-light. J. Food Sci. 2007, 72, M391–M396. [Google Scholar] [CrossRef] [PubMed]
  34. Ukuku, D.O.; Geveke, D.J. A combined treatment of UV-light and radio frequency electric field for the inactivation of Escherichia coli K-12 in apple juice. Int. J. Food Microbiol. 2010, 138, 50–55. [Google Scholar] [CrossRef] [PubMed]
  35. Cao, C.; Zhang, G.; Ye, J.; Hua, R.; Sun, Z.; Cui, J. Fast growth of self-aligned titania nanotube arrays with excellent transient photoelectric responses. Electrochim. Acta 2016, 211, 552–560. [Google Scholar] [CrossRef]
  36. Choi, H.; Al-Abed, S.R.; Agarwal, S.; Dionysiou, D.D. Synthesis of reactive nano-Fe/Pd bimetallic system-impregnated activated carbon for the simultaneous adsorption and dechlorination of PCBs. Chem. Mater. 2008, 20, 3649–3655. [Google Scholar] [CrossRef]
  37. Sharma, B.; Boruah, P.K.; Yadav, A.; Das, M.R. TiO2–Fe2O3 nanocomposite heterojunction for superior charge separation and the photocatalytic inactivation of pathogenic bacteria in water under direct sunlight irradiation. J. Environ. Chem. Eng. 2018, 6, 134–145. [Google Scholar] [CrossRef]
  38. Solano, R.A.; Herrera, A.P.; Maestre, D.; Cremades, A. Fe-TiO2 nanoparticles synthesized by green chemistry for potential application in waste water photocatalytic treatment. J. Nanotechnol. 2019, 2019, 4571848. [Google Scholar] [CrossRef] [Green Version]
  39. Zhu, J.; Chen, F.; Zhang, J.; Chen, H.; Anpo, M. Fe3+-TiO2 photocatalysts prepared by combining sol–gel method with hydrothermal treatment and their characterization. J. Photochem. Photobiol. A Chem. 2006, 180, 196–204. [Google Scholar] [CrossRef]
  40. Huerta-Flores, A.M.; Chávez-Angulo, G.; Carrasco-Jaim, O.A.; Torres-Martínez, L.M.; Garza-Navarro, M. Enhanced photoelectrochemical water splitting on heterostructured α-Fe2O3-TiO2: X (X = Co, Cu, Bi) photoanodes: Role of metal doping on charge carrier dynamics improvement. J. Photochem. Photobiol. A Chem. 2021, 410, 113077. [Google Scholar] [CrossRef]
  41. Wang, X.; Wang, A.; Lu, M.; Ma, J. Synthesis of magnetically recoverable Fe0/graphene-TiO2 nanowires composite for both reduction and photocatalytic oxidation of metronidazole. Chem. Eng. J. 2018, 337, 372–384. [Google Scholar] [CrossRef]
  42. Hsieh, W.-P.; Pan, J.R.; Huang, C.; Su, Y.-C.; Juang, Y.-J. Enhance the photocatalytic activity for the degradation of organic contaminants in water by incorporating TiO2 with zero-valent iron. Sci. Total. Environ. 2010, 408, 672–679. [Google Scholar] [CrossRef] [PubMed]
  43. Xia, H.; Xiong, W.; Lim, C.K.; Yao, Q.; Wang, Y.; Xie, J. Hierarchical TiO2-B nanowire@ α-Fe2O3 nanothorn core-branch arrays as superior electrodes for lithium-ion microbatteries. Nano Res. 2014, 7, 1797–1808. [Google Scholar] [CrossRef]
  44. Zhu, X.; Xu, H.; Yao, Y.; Liu, H.; Wang, J.; Pu, Y.; Feng, W.; Chen, S. Effects of Ag 0-modification and Fe3+-doping on the structural, optical and photocatalytic properties of TiO2. RSC Adv. 2019, 9, 40003–40012. [Google Scholar] [CrossRef] [Green Version]
  45. Tauc, J. Absorption edge and internal electric fields in amorphous semiconductors. Mater. Res. Bull. 1970, 5, 721–729. [Google Scholar] [CrossRef]
  46. Arellano, U.; Asomoza, M.; Ramirez, F. Antimicrobial activity of Fe–TiO2 thin film photocatalysts. J. Photochem. Photobiol. A Chem. 2011, 222, 159–165. [Google Scholar] [CrossRef]
  47. Zeng, G.; Zhang, Q.; Liu, Y.; Zhang, S.; Guo, J. Preparation of TiO2 and Fe-TiO2 with an impinging stream-rotating packed bed by the precipitation method for the photodegradation of gaseous toluene. Nanomaterials 2019, 9, 1173. [Google Scholar] [CrossRef] [Green Version]
  48. Chakhari, W.; Naceur, J.B.; Taieb, S.B.; Assaker, I.B.; Chtourou, R. Fe-doped TiO2 nanorods with enhanced electrochemical properties as efficient photoanode materials. J. Alloy. Compd. 2017, 708, 862–870. [Google Scholar] [CrossRef]
  49. Liu, Q.-P. Analysis on dye-sensitized solar cells based on Fe-doped TiO2 by intensity-modulated photocurrent spectroscopy and Mott–Schottky. Chin. Chem. Lett. 2014, 25, 953–956. [Google Scholar] [CrossRef]
  50. Tang, H.; Prasad, K.; Sanjines, R.; Schmid, P.; Levy, F. Electrical and optical properties of TiO2 anatase thin films. J. Appl. Phys. 1994, 75, 2042–2047. [Google Scholar] [CrossRef]
  51. Cui, H.; Zhao, W.; Yang, C.; Yin, H.; Lin, T.; Shan, Y.; Xie, Y.; Gu, H.; Huang, F. Black TiO2 nanotube arrays for high-efficiency photoelectrochemical water-splitting. J. Mater. Chem. A 2014, 2, 8612–8616. [Google Scholar] [CrossRef]
  52. Sahoo, P.P.; Zoellner, B.; Maggard, P.A. Optical, electronic, and photoelectrochemical properties of the p-type Cu3−xVO4 semiconductor. J. Mater. Chem. A 2015, 3, 4501–4509. [Google Scholar] [CrossRef]
  53. Seger, B.; Tilley, S.D.; Pedersen, T.; Vesborg, P.C.; Hansen, O.; Grätzel, M.; Chorkendorff, I. Silicon protected with atomic layer deposited TiO2: Conducting versus tunnelling through TiO2. J. Mater. Chem. A 2013, 1, 15089–15094. [Google Scholar] [CrossRef]
  54. Freitas, R.; Santanna, M.; Pereira, E. Preparation and characterization of TiO2 nanotube arrays in ionic liquid for water splitting. Electrochim. Acta 2014, 136, 404–411. [Google Scholar] [CrossRef]
  55. Kormann, C.; Bahnemann, D.W.; Hoffmann, M.R. Preparation and characterization of quantum-size titanium dioxide. J. Phys. Chem. 1988, 92, 5196–5201. [Google Scholar] [CrossRef]
  56. Ma, B.; Yu, N.; Xin, S.; Xin, Y.; Zhang, C.; Ma, X.; Gao, M. Photoelectrocatalytic degradation of p-chloronitrobenzene by g-C3N4/TiO2 nanotube arrays photoelectrodes under visible light irradiation. Chemosphere 2021, 267, 129242. [Google Scholar] [CrossRef]
Figure 1. SEM images for (a) TNAs (×100 k), (b) cross-section image of TNAs (×50 k), (c) Fe/TNAs-0.2 (×200 k), and (d) Fe/TNAs-0.5 (×200 k) (parentheses shows the magnification).
Figure 1. SEM images for (a) TNAs (×100 k), (b) cross-section image of TNAs (×50 k), (c) Fe/TNAs-0.2 (×200 k), and (d) Fe/TNAs-0.5 (×200 k) (parentheses shows the magnification).
Nanomaterials 11 01944 g001
Figure 2. The XRD patterns for TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5.
Figure 2. The XRD patterns for TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5.
Nanomaterials 11 01944 g002
Figure 3. XPS spectra of (a) XPS spectrum of Fe/TNAs-0.5, (b) C 1s, (c) Fe 2p, (d) O 1s, and (e) Ti 2p.
Figure 3. XPS spectra of (a) XPS spectrum of Fe/TNAs-0.5, (b) C 1s, (c) Fe 2p, (d) O 1s, and (e) Ti 2p.
Nanomaterials 11 01944 g003aNanomaterials 11 01944 g003b
Figure 4. UV-vis DRS spectra of TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5; insert plot: bandgap energy for TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5.
Figure 4. UV-vis DRS spectra of TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5; insert plot: bandgap energy for TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5.
Nanomaterials 11 01944 g004
Figure 5. Photocurrent-time response of TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5 (light source: 100 W hg lamp, applied potential: +1.0 V (vs. Ag/AgCl), electrolyte: 0.1 M NaCl, reference electrode: Ag/AgCl, counter electrode: Pt wire, and working electrode: TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5).
Figure 5. Photocurrent-time response of TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5 (light source: 100 W hg lamp, applied potential: +1.0 V (vs. Ag/AgCl), electrolyte: 0.1 M NaCl, reference electrode: Ag/AgCl, counter electrode: Pt wire, and working electrode: TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5).
Nanomaterials 11 01944 g005
Figure 6. Nyquist plot (a) and Bode plot (b) based on EIS analyses of TNA, Fe/TNAs-0.2, and Fe/TNAs-0.5.
Figure 6. Nyquist plot (a) and Bode plot (b) based on EIS analyses of TNA, Fe/TNAs-0.2, and Fe/TNAs-0.5.
Nanomaterials 11 01944 g006
Figure 7. Mott-Schottky plot of TNAs and Fe/TNAs at frequency of 100 Hz (working electrode: TNAs (or Fe/TNAs); counter electrode: Pt wire; reference electrode: Ag/AgCl; temperature: 298 K; electrolyte: 0.1 M NaCl; light source: 100 W Hg lamp; amplitude: 10 mV).
Figure 7. Mott-Schottky plot of TNAs and Fe/TNAs at frequency of 100 Hz (working electrode: TNAs (or Fe/TNAs); counter electrode: Pt wire; reference electrode: Ag/AgCl; temperature: 298 K; electrolyte: 0.1 M NaCl; light source: 100 W Hg lamp; amplitude: 10 mV).
Nanomaterials 11 01944 g007
Figure 8. Comparison of methyl orange degradation in PEC, photocatalytic (PC), electrochemical (EC) and photolysis (p) processes by (a)TNAs, (b) Fe/TNAs-0.2, and (c) Fe/TNAs-0.5 (light source: 100 W hg lamp, applied potential: +1.0 V (vs. Ag/AgCl), electrolyte: 10 ppm methyl orange in 0.1 M NaCl solution, reference electrode: Ag/AgCl, counter electrode: Pt wire, and working electrode: TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5).
Figure 8. Comparison of methyl orange degradation in PEC, photocatalytic (PC), electrochemical (EC) and photolysis (p) processes by (a)TNAs, (b) Fe/TNAs-0.2, and (c) Fe/TNAs-0.5 (light source: 100 W hg lamp, applied potential: +1.0 V (vs. Ag/AgCl), electrolyte: 10 ppm methyl orange in 0.1 M NaCl solution, reference electrode: Ag/AgCl, counter electrode: Pt wire, and working electrode: TNAs, Fe/TNAs-0.2, and Fe/TNAs-0.5).
Nanomaterials 11 01944 g008
Figure 9. PEC and P removal of E. coli (light source: 100 W hg lamp, applied potential: +1.0 V (vs. Ag/AgCl), electrolyte: E. coli in 0.1 M NaCl solution, reference electrode: Ag/AgCl, counter electrode: Pt wire, and working electrode: Fe/TNAs-0.5).
Figure 9. PEC and P removal of E. coli (light source: 100 W hg lamp, applied potential: +1.0 V (vs. Ag/AgCl), electrolyte: E. coli in 0.1 M NaCl solution, reference electrode: Ag/AgCl, counter electrode: Pt wire, and working electrode: Fe/TNAs-0.5).
Nanomaterials 11 01944 g009
Table 1. EDX analyses for TNA, Fe/TNAs-0.2, and Fe/TNAs-0.5.
Table 1. EDX analyses for TNA, Fe/TNAs-0.2, and Fe/TNAs-0.5.
ElementsC
(Atomict %)
O
(Atomict %)
Ti
(Atomict %)
Fe
(Atomict %)
Total
Sample
TNAs2.0363.8234.150.00100.0
Fe/TNAs-0.21.9763.2533.840.93100.0
Fe/TNAs-0.59.9666.0816.517.44100.0
Table 2. Element composition of XPS analyses for TNA, Fe/TNAs-0.2, and Fe/TNAs-0.5.
Table 2. Element composition of XPS analyses for TNA, Fe/TNAs-0.2, and Fe/TNAs-0.5.
ElementsC
(Atomict %)
O
(Atomict %)
Ti
(Atomict %)
Fe
(Atomict %)
Sample
Fe/TNAs-0.250.945.130.9
Fe/TNAs-0.553.427.7162.9
Table 3. Fitting results of equivalent circuits of TNAs and Fe/TNAs.
Table 3. Fitting results of equivalent circuits of TNAs and Fe/TNAs.
Rs(Ω)Rp(Ω)CPE(Ω)fmazτt (ms)
TNAs38.93209.500.00570.5484290.3
Fe/TNAs-0.235.68143.330.00660.4489354.7
Fe/TNAs-0.534.56130.970.00560.3674433.3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, C.-H.; Peng, Y.-P.; Lin, M.-H.; Chang, K.-L.; Lin, Y.-C.; Sun, J. Iron Modified Titanate Nanotube Arrays for Photoelectrochemical Removal of E. coli. Nanomaterials 2021, 11, 1944. https://doi.org/10.3390/nano11081944

AMA Style

Chen C-H, Peng Y-P, Lin M-H, Chang K-L, Lin Y-C, Sun J. Iron Modified Titanate Nanotube Arrays for Photoelectrochemical Removal of E. coli. Nanomaterials. 2021; 11(8):1944. https://doi.org/10.3390/nano11081944

Chicago/Turabian Style

Chen, Chia-Hung, Yen-Ping Peng, Ming-Hsun Lin, Ken-Lin Chang, Yung-Chang Lin, and Jian Sun. 2021. "Iron Modified Titanate Nanotube Arrays for Photoelectrochemical Removal of E. coli" Nanomaterials 11, no. 8: 1944. https://doi.org/10.3390/nano11081944

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