Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

High chlorine evolution performance of electrochemically reduced TiO2 nanotube array coated with a thin RuO2 layer by the self-synthetic method

Teayoung Leea, Woonghee Leeb, Seongsoo Kima, Changha Leea, Kangwoo Chob, Choonsoo Kim*c and Jeyong Yoon*ad
aSchool of Chemical and Biological Engineering, Institute of Chemical Processes (ICP), Seoul National University, 1 Gwanak-ro, Gwanak-gu, Seoul 08826, Republic of Korea. E-mail: jeyong@snu.ac.kr
bDivision of Environmental Science & Engineering, POSTECH, 77 Chungam-ro, Nam-gu, Pohang 37673, Republic of Korea
cDepartment of Environmental Engineering, Institute of Energy/Environment Convergence Technologies, Kongju National University, 1223-24, Cheonan-daero, Cheonan-si 31080, Republic of Korea. E-mail: choonsoo@kongju.ac.kr
dKorea Environment Institute, 370 Sicheong-daero, Sejong-si 30147, Republic of Korea

Received 12th November 2020 , Accepted 15th February 2021

First published on 25th March 2021


Abstract

Recently, reduced TiO2 nanotube arrays via electrochemical self-doping (r-TiO2) are emerging as a good alternative to conventional dimensionally stable anodes (DSAs) due to their comparable performance and low-cost. However, compared with conventional DSAs, they suffer from poor stability, low current efficiency, and high energy consumption. Therefore, this study aims to advance the electrochemical performances in the chlorine evolution of r-TiO2 with a thin RuO2 layer coating on the nanotube structure (RuO2@r-TiO2). The RuO2 thin layer was successfully coated on the surface of r-TiO2. This was accomplished with a self-synthesized layer of ruthenium precursor originating from a spontaneous redox reaction between Ti3+ and metal ions on the r-TiO2 surface and thermal treatment. The thickness of the thin RuO2 layer was approximately 30 nm on the nanotube surface of RuO2@r-TiO2 without severe pore blocking. In chlorine production, RuO2@r-TiO2 exhibited higher current efficiency (∼81.0%) and lower energy consumption (∼3.0 W h g−1) than the r-TiO2 (current efficiency of ∼64.7% of and energy consumption of ∼5.2 W h g−1). In addition, the stability (ca. 22 h) was around 20-fold enhancement in RuO2@r-TiO2 compared with r-TiO2 (ca. 1.2 h). The results suggest a new route to provide a thin layer coating on r-TiO2 and to synthesize a high performance oxidant-generating anode.


1. Introduction

The electrochemical oxidation process (EOP) has emerged as an alternative to the conventional oxidation process because of its relatively simple facilities, maintenance, and accessibility.1–8 In addition, as the need for a decentralized water treatment system has increased, the spectrum of EOP's application is broadening from urban to rural areas and developing countries.2,4,9–11 The process controls contaminants by generating oxidants on-site. In particular, chlorine (Cl2) has been widely used to remove microorganisms, organic matters, and ammonia effectively along with other oxidants such as ozone, hydroxyl, and sulphate radicals.1,4,5,10,12–16 The generated Cl2 diffuses to the bulk solution (below eqn (1)–(3)), and they exist three major species including Cl2 (pH < 3), HOCl (pH 3–8), and OCl (pH > 8).1
 
2Cl → Cl2(aq) + 2e (E0 = 1.36 V vs. NHE) (1)
 
Cl2(aq) + H2O → HOCl + Cl + H+ (2)
 
HOCl ↔ OCl + H+ (3)

Moreover, Cl2 is considered as a disinfectant to protect from the infectious COVID-19 in the water, sanitation, and hygiene (WASH) field by World Health Organization (WHO),17 and several researchers suggested using it for treating the wastewater from hospital or household against potentially dangerous coronavirus.18,19 Thus, practically, the Cl2 generation system by EOP has a lot of attention with the advantages and is thought to be a suitable technology for small communities to overcome the disease.

For the high efficiency of EOP, the anode material is a pivotal factor governing oxidant species, energy consumption, and cost-effectiveness. A dimensionally stable anode (DSA; RuO2, IrO2, etc.) has an excellent electrochemical property for chlorine evolution reaction (ClER),5,14,20–27 but, one obstacle in the effective use of DSA is the high manufacturing cost based on the inclusion of expensive noble metals.

In this regard, recently, a reduced TiO2 nanotube array (r-TiO2), which can be simply fabricated by electrochemical self-doping of an anatase TiO2 nanotube array (a-TiO2), has attracted much attentions as a promising electrode in electrochemical ClER.11,28–30 The self-doping simply converts Ti4+ in a-TiO2 to Ti3+ via the intercalation of protons as self-dopants, potentially leading to high electrocatalytic activity in the generation of oxidants with high surface area and low-cost.28,31 In spite of its advantages, r-TiO2 has unfortunately suffered from a poor long-term service time, low current efficiency and high energy consumption in chlorine production.

Therefore, this study aimed to enhance the chlorine generation performance of r-TiO2 with a simple thin layer coating of RuO2 as an excellent anode material (RuO2@r-TiO2). This RuO2 thin layer was coated on the surface of r-TiO2 via the spontaneous reduction of ruthenium precursor resulting from a partial conversion reaction of Ti3+ to Ti4+ (ref. 32 and 33) in r-TiO2 and the followed thermal treatment, successfully leading to the improvement of electrocatalyst for chlorine evolution. To fully understand its surface properties, we used field emission scanning electron microscopy (FE-SEM), field emission transmission electron microscopy (FE-TEM), X-ray diffraction (XRD) and X-ray photoelectron spectroscopy (XPS). The electrocatalytic activity for ClER was investigated with cyclic voltammetry (CV) measurement and the N,N-diethyl-p-phenylenediamine (DPD) method. Furthermore, scanning electrochemical microscopy (SECM) was employed as an in situ analysis to investigate the enhanced ClER from the uniformly coated RuO2 on the r-TiO2.

2. Materials and methods

2.1 Preparation of r-TiO2 and RuO2@TiO2

r-TiO2 was prepared by a typical two-step anodization method and electrochemical self-doping. First, the Ti foil was anodized at 60 V for 2 h in an ethylene-glycol-based electrolyte containing DI water (2.5 wt%) and NH4F (0.2 wt%). The formed nanotube film was peeled off by a compressed air stream, and then a second-anodization was conducted at 40 V for 7 h under the same electrolyte condition. By annealing the as-prepared TiO2 NTA at 450 °C for 1 h in air, the crystal structure was converted into an anatase-dominant phase (a-TiO2).34 Then, electrochemical doping was performed on the a-TiO2 under cathodic polarization with constant current (16.7 mA cm−2) for 90 s in a phosphate buffer solution ([KH2PO4]0 = 0.1 M with KOH, pH = 7.02).28–30 The prepared r-TiO2 had a thickness of approximately 13.8 μm and the width of the nanotubes was 130 ± 30 nm (Fig. S1).

Prior to the RuO2 coating process, the r-TiO2 was fully dried at room temperature, then immersed in an aqueous ruthenium precursor solution (5 mM RuCl3·H2O in deionized (DI) water) under the dark condition for 24 h to produce the self-synthesized coating (Ru@r-TiO2). After that, the treated r-TiO2 was fully washed with DI water to remove the remained ruthenium precursor in the nanotubes, then annealed in 450 °C (air) for 1 h to be oxidized (RuO2@r-TiO2). The process is presented briefly in Fig. 1.


image file: d0ra09623g-f1.tif
Fig. 1 Scheme of thin RuO2 layer coating process on reduced TiO2 nanotube array via electrochemical self-doping.

2.2 Characterization of RuO2@r-TiO2

The morphologies of r-TiO2 and RuO2@r-TiO2 were observed with field emission scanning electron microscopy (FE-SEM, JSM-6701F, JEOL, Japan) at 20 kV, and field emission transmission electron microscopy (FE-TEM, JEM-F200, JEOL, Japan) was employed to confirm the deposited RuO2 layer on the wall of r-TiO2 including energy-dispersive X-ray spectroscopy (EDS). An X-ray diffractometer (XRD, Bruker D8 DISCOVER, Germany) and X-ray photoelectron spectroscope (XPS, Sigma Probe, ThermoVG, UK) were used to examine the material species of the TiO2 NTA-based electrodes.

2.3 Evaluation of electrochemical properties of RuO2@TiO2

The electrochemical properties of RuO2@r-TiO2 were investigated by cyclic voltammetry (CV) measurement with a three-electrodes system (reference electrode: Ag/AgCl in sat. KCl, counter electrode: Pt mesh) at a scan rate of 5 mV s−1. To understand the electrocatalytic activity of RuO2@r-TiO2, chlorine was electrochemically produced in a two-electrodes system that consisted of RuO2@r-TiO2 as an anode and Pt mesh as a cathode with a constant current density of 16.7 mA cm−2 in 0.1 M NaCl. The produced chlorine concentration was monitored by the N,N-diethyl-p-phenylenediamine (DPD) method with a spectrophotometer (DR 900, Hach Co., USA, 530 nm). The current efficiency (%) and the energy consumption (W h g−1) of chlorine generation were calculated by eqn (4) and (5).
 
image file: d0ra09623g-t1.tif(4)
 
image file: d0ra09623g-t2.tif(5)
where c is the concentration of generated chlorine (g L−1), V is the electrolyte volume (L), n is the electrons' number (1 eq. mol−1), F is the faradaic constant (96[thin space (1/6-em)]485 C eq.−1), M(Cl2) is the molecular weight of chlorine (71 g mol−1), I is the applied constant current (C s−1), t is the operation time (s), and e is the cell voltage (V).

To confirm the chlorine evolution mechanism of RuO2@r-TiO2, the hydroxyl radical production was investigated with a degradation of terephthalic acid (TA) as a probe compound. In addition, the effect of hydroxyl radical on chlorine evolution was examined by adding 1 M t-BuOH. The TA degradation was measured by high-performance liquid chromatography (HPLC; Ultimate 3000, Dionex, Sunnyvale, CA, USA) in the methanol and formic acid (0.1%) mixture eluent (v/v, 60[thin space (1/6-em)]:[thin space (1/6-em)]40).35 Furthermore, scanning electrochemical microscopy [SECM; SP-300 (bipotentiostat), M470 (SECM Workstation, Bio-Logics SAS), France] was performed to scrutinize the activities for ClER on samples as an in situ measurement. This visualized the scanned area (500 μm × 500 μm of the electrodes) with a colour gradation from blue to red.

3. Results and discussion

3.1 Morphology of RuO2@r-TiO2

Fig. 2 shows FE-SEM images (top and cross sectional view) of r-TiO2 and RuO2@r-TiO2. As can be seen in Fig. 2, significant differences on the nanotube edge and sidewall of RuO2@r-TiO2 were found. The RuO2@r-TiO2 revealed that a large number of nano-grains were formed on the verges of the nanopore entrances compared to the r-TiO2 in Fig. 2a and c. In addition, the sidewall of RuO2@r-TiO2 was thicker after deposition on r-TiO2 (Fig. 2b and d). It seems that the assembles of nano-grains on the sidewall of the inner pore were formed to a thin layer (see inset image of Fig. 2d).
image file: d0ra09623g-f2.tif
Fig. 2 Morphologies of (a) top- (b) cross sectional view of r-TiO2 and (c) top- (d) cross sectional view of RuO2 @r-TiO2 with a field emission scanning electron microscope (FE-SEM).

The sidewall condition of RuO2@r-TiO2 is inspected meticulously by FE-TEM in Fig. 3. Considering the thickness of the r-TiO2 sidewall (ca. 14–22 nm), the sidewall thickness of RuO2@r-TiO2 (ca. ∼45.3 nm) was approximately three times thicker. With the distinguished interface layer at RuO2@r-TiO2 (red dashed square in inset image of Fig. 3b) and the results of EDS in Fig. S2, we see that RuO2 was successfully deposited on the sidewall of RuO2@r-TiO2. It is plausible that the RuO2@r-TiO2 revealed a fine coating on the overall surface of r-TiO2 with open-top and hollow nanotube structure via the self-synthesized coating method.


image file: d0ra09623g-f3.tif
Fig. 3 Nanotube structure characterization of (a) r-TiO2 and (b) RuO2@r-TiO2 using field emission transmission electron microscope (FE-TEM).

3.2 Characteristics of the deposits on RuO2@r-TiO2

To better understand the deposits on RuO2@r-TiO2, the XRD patterns and XPS spectra of as-prepared r-TiO2 and RuO2@r-TiO2 are further analysed in Fig. 4. From the results of the XRD patterns in Fig. 4a, an elusive peak appears at 28° with regard to RuO2 at RuO2@r-TiO2.36–38 Hence, the coated RuO2 is examined in detail through the XPS results of Ru 3d, O 1s, and Ti 2p in Fig. 4b–d. As shown in Fig. 4b, the RuO2@r-TiO2 exhibited a clear RuO2 peak at 280.4 eV in the XPS spectra of Ru 3d5/2.38–43 From the shoulder peak (529.4 eV) in XPS spectra of O 1s (red line in Fig. 4c), the deposition of RuO2 was further confirmed.38,39,44–46 Moreover, this is clearly supported by the peak shift (1.2 eV) from 458.9 to 457.7 eV in Fig. 4d of Ti 2p3/2 indicating a heterojunction of RuO2 and TiO2.40,42,47
image file: d0ra09623g-f4.tif
Fig. 4 (a) XRD patterns and XPS spectra (b) Ru 3d, (c) O 1s, and (d) Ti 2p of r-TiO2 and RuO2@r-TiO2.

These results suggest that the RuO2 thin layer was well-formed on the overall surface of RuO2@r-TiO2 via the self-synthesized coating method. Note that the self-synthesis method has been reported in the nanoparticles of noble metals on a TiO2 sphere with the spontaneous redox reaction between metal ions and the reduced TiO2.32,33 To the best of our knowledge, this is the first report to prepare a thin RuO2 layer coating on electrochemically reduced TiO2 NTA via the previous phenomenon including the following thermal treatment, without severe pore blockage and toxic chemicals when compared to other methods for treating active materials on the TiO2 NTAs.37,48–53

3.3 Electrocatalytic activities of RuO2@r-TiO2

Fig. 5 shows the improved electrochemical properties of RuO2@r-TiO2. From the result in Fig. 5a, where the RuO2@r-TIO2 initiated an oxygen evolution reaction (OER) at a potential of approximately 1.1 V vs. Ag/AgCl, the over-potential of which decreased significantly by up to 1.0 V compared to the r-TiO2. This means that the surface of RuO2@r-TiO2 had a higher electrocatalytic activity for OER when assisted by the thin RuO2 layer. In the cathodic biased potential regime ranging from 0 to −1.5 V vs. Ag/AgCl on the RuO2@r-TiO2, there were no reactions regarding proton inter/deintercalation (−0.6/−0.9 V vs. Ag/AgCl) which was obviously found on r-TiO2 (blue dash line in the inset image of Fig. 5a) as a unique electrochemical feature of r-TiO2.28,54–57 Additionally, as shown in Fig. S3, in contrast to RuO2@r-TiO2, the r-TiO2 lost its electrochemical property after the thermal treatment. This implies that the self-synthesized coating covered the entire surface of the r-TiO2 and that the r-TiO2 under the deposit of RuO2@r-TiO2 was prevented from oxidizing during the thermal treatment. Accordingly, its electrochemical property did not vanish even when it was annealed in the air condition; rather, this property was improved by the formed RuO2.
image file: d0ra09623g-f5.tif
Fig. 5 (a) Cyclic voltammograms (CV) (scan rate: 5 mV s−1), the stability test with (b) applied constant current density (16.7 mA cm−2), and (c) polarity reversal operation (±16.7 mA cm−2, switching time 90 s) of r-TiO2 and RuO2@r-TiO2 in 0.1 M phosphate buffer solution (PBS).

For oxygen evolution under a constant current condition (Fig. 5b), RuO2@r-TiO2 led to a lower initial operational cell voltage (∼3.0 V) than r-TiO2 (∼4.2 V), and it showed significantly enhanced stability with a value that was approximately 20 times higher (∼22 h) than that of r-TiO2 (∼1.2 h). This can be evaluated based on the drastic increase in cell voltage. In addition, under a polarity reversal operation (switching constant current for +16.7 mA cm−2 vs. −16.7 mA cm−2), the RuO2@r-TiO2 was highly stable on the stress from the harsh reversal condition compared to r-TiO2. This indicates that RuO2@r-TiO2 is a more reliable material than r-TiO2 in various environmental and industrial applications. Nevertheless, for the further success of RuO2@r-TiO2, the long-term performance stability must be improved via controlling the doping level of r-TiO2 (Fig. S4 in ESI). It is required to further study the effect of type of catalyst (i.e., IrO2, Pt and carbon, etc.) for thin layer coating, coating thickness, and temperature of thermal treatment on CIER.

Fig. 6 shows the enhanced chlorine evolution performance and the pathway for ClER of RuO2@r-TiO2 compared to r-TiO2. As shown in Fig. 6a, the chlorine production rate of RuO2@r-TiO2 was estimated to be approximately 17.85 mg L−1 min−1. This is approximately 20% faster than that of r-TiO2 (14.35 mg L−1 min−1). Correspondingly, RuO2@r-TiO2 exhibited a current efficiency of 81.0% with an energy consumption of 3.0 W h g−1, indicating higher electrocatalytic activity for chlorine production compared to r-TiO2 (current efficiency of 64.7% and energy consumption of 5.2 W h g−1 in Fig. 6b). The high chlorine evolution performances can be explained by the uniformly organized nanotube structure with the thin RuO2 layer.


image file: d0ra09623g-f6.tif
Fig. 6 (a) Evolution of chlorine (16.7 mA cm−2, 0.1 M NaCl) and (b) the current efficiency and energy consumption in 3 min, and (c) terephthalic acid (TA) degradation for hydroxyl radical measurement (0.1 mM TA, 16.7 mA cm−2) of RuO2@r-TiO2 and r-TiO2.

Furthermore, with the thin RuO2 layer, the surface of RuO2@r-TiO2 was converted to be more attractive for chlorine than hydroxyl radical (Fig. 6c). This resulted in excellent chlorine production performances, namely the RuO2@r-TiO2 can be defined as an active electrode (high efficiency for chlorine production; RuO2, IrO2, etc.) rather than an inactive electrode (high efficiency for hydroxyl radical; r-TiO2, boron doped diamond electrode, SnO2, PbO2, etc.).5,14,20,28,37 Note that, in common, the active electrode produces CIER via direct electron transfer with chloride ions whereas the inactive electrode leads to CIER by the indirect pathway mediated by hydroxyl radicals. This is well supported by the effect of the hydroxyl radical scavenger on the chlorine evolution (Fig. S5). With the addition of t-BuOH as hydroxyl radical scavenger, the chlorine evolution efficiency of RuO2@r-TiO2 did not meaningfully decrease, whereas that of r-TiO2 was significantly reduced. This means the small effect of hydroxyl radical on the chlorine evolution of RuO2@r-TiO2 instead of r-TiO2 produced chlorine mediated by hydroxyl radical.39 Considering the effect of hydroxyl radical on CIER of RuO2@r-TiO2 and r-TiO2, the surface of RuO2@r-TiO2 behaves active electrodes. Eventually, we see the surface of RuO2@r-TiO2 was converted to an active electrode from an inactive electrode. It is attributed to that the RuO2 thin layer was uniformly coated on the surface of nanopores of r-TiO2, and thus, the surface of r-TiO2 only worked as substrate, not catalytic material.

3.4 Effect of the immersion time on chlorine production efficiency of RuO2@r-TiO2

To optimize the deposition of RuO2 on RuO2@r-TiO2, we further investigate the effect of the RuO2 loading amount controlled via Ru precursor dipping time (ranging from 1 to 24 h) on the deposition and chlorine evolution efficiency of RuO2@r-TiO2. Fig. 7 shows the morphologies of the RuO2@r-TiO2 samples prepared with the time of 1, 3, 6, 12, and 24 h. As shown in Fig. 7, as the immersion time increased, the size of the nano-grains on the nano-pore edge and the thickness of the sidewall (inset images) gradually increased. In particular, regardless of the time, surfaces of all samples were more attractive for chlorine evolution than hydroxyl radical (Fig. 8a). However, compared to the chlorine evolution performances of RuO2@r-TiO2 prepared with the immersion times of 6, 12 and 24 h (current efficiency of 79.2, 79.4, and 81.0%; and energy consumption of 3.0, 3.0, 3.0 W h g−1, respectively), the RuO2@r-TiO2 prepared with immersion times of 1 and 3 h resulted in relatively low chlorine evolution performance with current efficiencies of 72.2, 77.6% and energy consumption of 3.6, 3.2 W h g−1, respectively (see Fig. 8b, and refer to all data of chlorine generation in Fig. S6a). Correspondingly, a similar trend in the long-term stability was found (Fig. S6b).
image file: d0ra09623g-f7.tif
Fig. 7 FE-SEM images (cross sectional view in inset images) of (a) r-TiO2 and RuO2 coated electrodes after immersion in 5 mM aqueous ruthenium precursor for (b) 1, (c) 3, (d) 6, (e) 12, and (f) 24 h. Note that FE-SEM images of r-TiO2 and RuO2@r-TiO2 (24 h) were from the result of Fig. 2 in order to compare the morphologies of each electrode.

image file: d0ra09623g-f8.tif
Fig. 8 (a) OH radical generation measurement with TA degradation (0.1 mM TA, 16.7 mA cm−2) and (b) chlorine generation efficiency and energy consumption (16.7 mA cm−2, 0.1 M NaCl, 3 min) of r-TiO2 and RuO2 coated electrodes after immersion in 5 mM aqueous ruthenium precursor for 1, 3, 6, 12, and 24 h. Note that chlorine generation efficiency, energy consumption, and the TA degradation data of r-TiO2, and RuO2@r-TiO2 (24 h) are from the results of Fig. 6.

Moreover, this is well supported by the results obtained with the sample generation/tip collection (SG/TC) mode of SECM (Fig. 9) which visualized the electrocatalytic activity for chlorine production via a chlorine reduction reaction at −0.2 V vs. Ag/AgCl (the detailed experimental condition are shown in Fig. S7 and S8).58,59 As shown in Fig. 9, the electrocatalytic activity was evenly enhanced across the entire surface after 12 h according to the condition of deposited RuO2 confirmed previously in Fig. 7 with FE-SEM. Particularly, compared to the pristine r-TiO2 (Fig. 9a) which revealed nano-patterns of emerald-colored rings, reddish cores were observed after commencing RuO2 deposition on r-TiO2 (Fig. 9b–f). The RuO2 coated inner pores improved the chlorine generation inside of the nanotubes, so that the higher activity (red) was measured at the center of the patterns than the circular edge. In Fig. 9f, circular nano-patterns on RuO2@r-TiO2 (24 h) were uniformly distributed in red. This color change verifies that the RuO2 was effectively coated at the top edge of the tubes and even at its inner-pores as previously shown in images of SEM and TEM (Fig. 2 and 3), resulting in the enhanced the electrocatalytic performance for the CIER. As such, the morphological and electrochemical properties of r-TiO2 are feasible to be facilely controlled with the self-synthesis coating method that can be extended to various fields of studies utilizing TiO2 NTA-based electrocatalysts.


image file: d0ra09623g-f9.tif
Fig. 9 Area scan of scanning electrochemical microscopy (SECM) in 0.1 M NaCl, 16.7 mA cm−2; (a) r-TiO2, RuO2 coated electrodes after immersing in 5 mM aqueous ruthenium precursor for (b) 1, (c) 3, (d) 6, (e) 12, and (f) 24 h.

4. Conclusions

In this study, we successfully fabricated the RuO2@r-TiO2 with a self-synthesized coating method leading to a RuO2 thin layer coating on r-TiO2 and demonstrated considerably enhanced electro-catalytic activity for chlorine production. The fine RuO2 coating was achieved via the self-synthesized coating originating from the conversion of Ti3+ to Ti4+ on r-TiO2 and thermal treatment under the atmospheric condition. Using various surface analysis including FE-SEM, FE-TEM, EDS and XPS, the formation of a RuO2 thin layer (thickness of ∼27.5 nm) on the inner pore sidewall of RuO2@r-TiO2 was clearly proven. The RuO2@r-TiO2 exhibited the highly enhanced electrocatalytic activity for chlorine production with the production rate of 17.85 mg L−1 min−1, the high current efficiency of 81.0%, energy consumption of 3.0 W h g−1, and long-term stability of ∼22 h compared to r-TiO2 (production rate of 14.35 mg L−1 min−1, the current efficiency of 64.7%, energy consumption of 5.2 W h g−1, and long term service time of 1.2 h). In addition, the performances of RuO2@r-TiO2 was optimized by controlling the immersion time in the precursor. These results provide a new approach to the thin metal oxide coating on r-TiO2 and provide opportunities for various applications such as electrolysis, photo-catalyst, and energy storage devices.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This research was supported by the Technology Innovation Program (10082572, Development of Low Energy Desalination Water Treatment Engineering Package System for Industrial Recycle Water Production) funded by the Ministry of Trade, Industry & Energy (MOTIE, Korea) and the National Research Foundation of Korea (NRF) grant funded by the Ministry of Science and ICT of the Korea Government (MSIT) (NRF-2019R1G1A1003336).

References

  1. Z. Chen, Y. Liu, W. Wei and B. Ni, Recent advances in electrocatalysts for halogenated organic pollutant degradation, Environ. Sci.: Nano, 2019, 6, 2332 RSC.
  2. S. O. Ganiyu, C. A. Martínez-Huitle and M. A. Rodrigo, Renewable energies driven electrochemical wastewater/soil decontamination technologies: a critical review of fundamental concepts and applications, Appl. Catal., B, 2020, 270, 118857 CrossRef CAS.
  3. V. Poza-Nogueiras, M. Pazos, M. Á. Sanromán and E. González-Romero, Double benefit of electrochemical techniques: treatment and electroanalysis for remediation of water polluted with organic compounds, Electrochim. Acta, 2019, 320, 1–13 CrossRef.
  4. B. C. Hodges, E. L. Cates and J. H. Kim, Challenges and prospects of advanced oxidation water treatment processes using catalytic nanomaterials, Nat. Nanotechnol., 2018, 13, 642–650 CrossRef CAS PubMed.
  5. M. Panizza and G. Cerisola, Direct and Mediated Anodic Oxidation of Organic Pollutants, Chem. Rev., 2009, 109, 6541–6569 CrossRef CAS PubMed.
  6. F. C. Moreira, R. A. R. Boaventura, E. Brillas and V. J. P. Vilar, Electrochemical advanced oxidation processes: a review on their application to synthetic and real wastewaters, Appl. Catal., B, 2017, 202, 217–261 CrossRef CAS.
  7. C. A. Martínez-Huitle and S. Ferro, Electrochemical oxidation of organic pollutants for the wastewater treatment: direct and indirect processes, Chem. Soc. Rev., 2006, 35, 1324–1340 RSC.
  8. J. Grimm, D. Bessarabov and R. Sanderson, Review of electro-assisted methods for water purification, Desalination, 1998, 115, 285–294 CrossRef CAS.
  9. K. Cho and M. R. Hoffmann, Urea degradation by electrochemically generated reactive chlorine species: products and reaction pathways, Environ. Sci. Technol., 2014, 48, 11504–11511 CrossRef CAS PubMed.
  10. X. Huang, Y. Qu, C. A. Cid, C. Finke, M. R. Hoffmann, K. Lim and S. C. Jiang, Electrochemical disinfection of toilet wastewater using wastewater electrolysis cell, Water Res., 2016, 92, 164–172 CrossRef CAS PubMed.
  11. Y. Yang and M. R. Hoffmann, Synthesis and Stabilization of Blue-Black TiO2 Nanotube Arrays for Electrochemical Oxidant Generation and Wastewater Treatment, Environ. Sci. Technol., 2016, 50, 11888–11894 CrossRef CAS PubMed.
  12. S. Trasatti, Electrochemistry and environment: the role of electrocatalysis, Int. J. Hydrogen Energy, 1995, 20, 835–844 CrossRef CAS.
  13. H. F. Diao, X. Y. Li, J. D. Gu, H. C. Shi and Z. M. Xie, Electron microscopic investigation of the bactericidal action of electrochemical disinfection in comparison with chlorination, ozonation and Fenton reaction, Process Biochem., 2004, 39, 1421–1426 CrossRef CAS.
  14. C. Comninellis, Electrocatalysis in the Electrochemical Conversion/Combustion of Organic Pollutants for Waste Water Treatment, Electrochim. Acta, 1994, 39, 1857–1862 CrossRef CAS.
  15. J. Kim, C. Lee and J. Yoon, Electrochemical Peroxodisulfate (PDS) Generation on a Self-Doped TiO2 Nanotube Array Electrode, Ind. Eng. Chem. Res., 2018, 57, 11465–11471 CrossRef CAS.
  16. J. Jeong, C. Kim and J. Yoon, The effect of electrode material on the generation of oxidants and microbial inactivation in the electrochemical disinfection processes, Water Res., 2009, 43, 895–901 CrossRef CAS PubMed.
  17. Water, sanitation, hygiene and waste management for SARS-CoV-2, the virus that causes COVID-19, World Health Organization, https://www.who.int/publications/i/item/WHO-2019-nCoV-IPC-WASH-2020.4, accessed Oct. 2020 Search PubMed.
  18. G. D. Bhowmick, D. Dhar, D. Nath, M. M. Ghangrekar, R. Banerjee, S. Das and J. Chatterjee, Coronavirus disease 2019 (COVID-19) outbreak: some serious consequences with urban and rural water cycle, npj Clean Water, 2020, 3, 32 CrossRef CAS.
  19. J. Wang, J. Shen, D. Ye, X. Yan, Y. Zhang, W. Yang, X. Li, J. Wang, L. Zhang and L. Pan, Disinfection technology of hospital wastes and wastewater: suggestions for disinfection strategy during coronavirus disease 2019 (COVID-19) pandemic in China, Environ. Pollut., 2020, 262, 114665 CrossRef CAS PubMed.
  20. S. Trasatti, Progress in the Understanding of the Mechanism of Chlorine Evolution at Oxide Electrodes, Electrochim. Acta, 1987, 32, 369–382 CrossRef CAS.
  21. C. Comninellis and G. P. Vercesi, Characterization of DSA-type electrodes: choice of a coating, J. Appl. Electrochem., 1991, 21, 335–345 CrossRef CAS.
  22. C. E. Finke, S. T. Omelchenko, J. T. Jasper, M. F. Lichterman, C. G. Read, N. S. Lewis and M. R. Hoffmann, Enhancing the activity of oxygen-evolution and chlorine-evolution electrocatalysts by atomic layer deposition of TiO2, Energy Environ. Sci., 2019, 12, 358–365 RSC.
  23. Y. Lee, J. Suntivich, K. J. May, E. E. Perry and Y. Shao-Horn, Synthesis and activities of rutile IrO2 and RuO2 nanoparticles for oxygen evolution in acid and alkaline solutions, J. Phys. Chem. Lett., 2012, 3, 399–404 CrossRef CAS PubMed.
  24. D. Dionisio, L. H. E. Santos, M. A. Rodrigo and A. J. Motheo, Electro-oxidation of methyl paraben on DSA®-Cl2: UV irradiation, mechanistic aspects and energy consumption, Electrochim. Acta, 2020, 338, 135901 CrossRef CAS.
  25. A. Cornell, B. Håkansson and G. Lindbergh, Ruthenium based DSA® in chlorate electrolysis - critical anode potential and reaction kinetics, Electrochim. Acta, 2003, 48, 473–481 CrossRef CAS.
  26. K. S. Exner, J. Anton, T. Jacob and H. Over, Controlling selectivity in the chlorine evolution reaction over RuO2-based catalysts, Angew. Chem., Int. Ed., 2014, 53, 11032–11035 CrossRef CAS PubMed.
  27. S. Trasatti, Electrocatalysis: understanding the success of DSA®, Electrochim. Acta, 2000, 45, 2377–2385 CrossRef CAS.
  28. C. Kim, S. Kim, J. Choi, J. Lee, J. S. Kang, Y. E. Sung, J. Lee, W. Choi and J. Yoon, Blue TiO2 nanotube array as an oxidant generating novel anode material fabricated by simple cathodic polarization, Electrochim. Acta, 2014, 141, 113–119 CrossRef CAS.
  29. C. Kim, S. Kim, J. Lee, J. Kim and J. Yoon, Capacitive and oxidant generating properties of black-colored TiO2 nanotube array fabricated by electrochemical self-doping, ACS Appl. Mater. Interfaces, 2015, 7, 7486–7491 CrossRef CAS PubMed.
  30. C. Kim, S. Kim, S. P. Hong, J. Lee and J. Yoon, Effect of doping level of colored TiO2 nanotube arrays fabricated by electrochemical self-doping on electrochemical properties, Phys. Chem. Chem. Phys., 2016, 18, 14370–14375 RSC.
  31. P. Roy, S. Berger and P. Schmuki, TiO2 Nanotubes: Synthesis and Applications, Angew. Chem., Int. Ed., 2011, 2904–2939 CrossRef CAS PubMed.
  32. Z. Zheng, B. Huang, X. Qin, X. Zhang, Y. Dai and M. H. Whangbo, Facile in situ synthesis of visible-light plasmonic photocatalysts M@TiO2 (M = Au, Pt, Ag) and evaluation of their photocatalytic oxidation of benzene to phenol, J. Mater. Chem., 2011, 21, 9079–9087 RSC.
  33. Y. Xie, K. Ding, Z. Liu, R. Tao, Z. Sun, H. Zhang and G. An, In situ controllable loading of ultrafine noble metal particles on titania, J. Am. Chem. Soc., 2009, 131, 6648–6649 CrossRef CAS PubMed.
  34. C. Kim, S. Lee, S. Kim and J. Yoon, Effect of Annealing Temperature on the Capacitive and Oxidant-generating Properties of an Electrochemically Reduced TiO2 Nanotube Array, Electrochim. Acta, 2016, 222, 1578–1584 CrossRef CAS.
  35. Y. Jing and B. P. Chaplin, Mechanistic Study of the Validity of Using Hydroxyl Radical Probes to Characterize Electrochemical Advanced Oxidation Processes, Environ. Sci. Technol., 2017, 51, 2355–2365 CrossRef CAS PubMed.
  36. M. Etzi Coller Pascuzzi, A. Goryachev, J. P. Hofmann and E. J. M. Hensen, Mn promotion of rutile TiO2-RuO2 anodes for water oxidation in acidic media, Appl. Catal., B, 2020, 261, 10 CrossRef.
  37. J. Kim, C. Kim, S. Kim and J. Yoon, RuO2 coated blue TiO2 nanotube array (blue TNA-RuO2) as an effective anode material in electrochemical chlorine generation, J. Ind. Eng. Chem., 2018, 66, 478–483 CrossRef CAS.
  38. Q. Gu, Z. Gao, S. Yu and C. Xue, Constructing Ru/TiO2 Heteronanostructures Toward Enhanced Photocatalytic Water Splitting via a RuO2/TiO2 Heterojunction and Ru/TiO2 Schottky Junction, Adv. Mater. Interfaces, 2016, 3, 17–21 Search PubMed.
  39. J. F. Moulder, W. F. Stickle, P. E. Sobol and K. D. Bomben, Handbook of X-ray Photoelectron Spectroscopy, Physical Electrocnis, Inc., Minesota, 1995 Search PubMed.
  40. Y. He, D. Langsdorf, L. Li and H. Over, Versatile model system for studying processes ranging from heterogeneous to photocatalysis: epitaxial RuO2(110) on TiO2(110), J. Phys. Chem. C, 2015, 119, 2692–2702 CrossRef CAS PubMed.
  41. D. J. Morgan, Resolving ruthenium: XPS studies of common ruthenium materials, Surf. Interface Anal., 2015, 47, 1072–1079 CrossRef CAS.
  42. M. T. Uddin, Y. Nicolas, C. Olivier, T. Toupance, M. M. Müller, H. J. Kleebe, K. Rachut, J. Ziegler, A. Klein and W. Jaegermann, Preparation of RuO2/TiO2 mesoporous heterostructures and rationalization of their enhanced photocatalytic properties by band alignment investigations, J. Phys. Chem. C, 2013, 117, 22098–22110 CrossRef CAS.
  43. H. Yue, L. Xue and F. Chen, Efficiently electrochemical removal of nitrite contamination with stable RuO2-TiO2/Ti electrodes, Appl. Catal., B, 2017, 206, 683–691 CrossRef CAS.
  44. D. D. Sarma and C. N. R. Rao, XPES studies of oxides of second- and third-row transition metals including rare earths, J. Electron Spectrosc. Relat. Phenom., 1980, 20, 25–45 CrossRef CAS.
  45. K. S. Kim and N. Winograd, X-ray Photoelectron Spectroscopic Studies of Ruthenium-Oxygen Surfaces, J. Catal., 1974, 72, 66–72 CrossRef.
  46. J. Y. Shen, A. Adnot and S. Kaliaguine, An ESCA study of the interaction of oxygen with the surface of ruthenium, Appl. Surf. Sci., 1991, 51, 47–60 CrossRef CAS.
  47. W. Ouyang, M. J. Muñoz-Batista, A. Kubacka, R. Luque and M. Fernández-García, Enhancing photocatalytic performance of TiO2 in H2 evolution via Ru co-catalyst deposition, Appl. Catal., B, 2018, 238, 434–443 CrossRef CAS.
  48. J. M. Macak, B. G. Gong, M. Hueppe and P. Schmuki, Filling of TiO2 nanotubes by self-doping and electrodeposition, Adv. Mater., 2007, 19, 3027–3031 CrossRef CAS.
  49. A. Gao, R. Hang, X. Huang, L. Zhao, X. Zhang, L. Wang, B. Tang, S. Ma and P. K. Chu, The effects of titania nanotubes with embedded silver oxide nanoparticles on bacteria and osteoblasts, Biomaterials, 2014, 35, 4223–4235 CrossRef CAS PubMed.
  50. J. A. Seabold, K. Shankar, R. H. T. Wilke, M. Paulose, O. K. Varghese, C. A. Grimes and K. S. Choi, Photoelectrochemical properties of heterojunction CdTe/TiO2 electrodes constructed using highly ordered TiO2 nanotube arrays, Chem. Mater., 2008, 20, 5266–5273 CrossRef CAS.
  51. H. Yoo, K. Oh, G. Lee and J. Choi, RuO2-Doped Anodic TiO2 Nanotubes for Water Oxidation: Single-Step Anodization vs. Potential Shock Method, J. Electrochem. Soc., 2017, 164, H104–H111 CrossRef CAS.
  52. N. Denisov, J. E. Yoo and P. Schmuki, Effect of different hole scavengers on the photoelectrochemical properties and photocatalytic hydrogen evolution performance of pristine and Pt-decorated TiO2 nanotubes, Electrochim. Acta, 2019, 319, 61–71 CrossRef CAS.
  53. J. Huang, M. Hou, J. Wang, X. Teng, Y. Niu, M. Xu and Z. Chen, RuO2 nanoparticles decorate belt-like anatase TiO2 for highly efficient chlorine evolution, Electrochim. Acta, 2020, 339, 1–9 CrossRef.
  54. S. P. Hong, S. Kim, N. Kim, J. Yoon and C. Kim, A short review on electrochemically self-doped TiO2 nanotube arrays: synthesis and applications, Korean J. Chem. Eng., 2019, 36, 1753–1766 CrossRef CAS.
  55. A. Ghicov, H. Tsuchiya, R. Hahn, J. M. MacAk, A. G. Muñoz and P. Schmuki, TiO2 nanotubes: H+ insertion and strong electrochromic effects, Electrochem. Commun., 2006, 8, 528–532 CrossRef CAS.
  56. H. Tokudome and M. Miyauchi, Electrochromism of titanate-based nanotubes, Angew. Chem., Int. Ed., 2005, 44, 1974–1977 CrossRef CAS PubMed.
  57. N. Sakai, A. Fujishima, T. Watanabe and K. Hashimoto, Highly Hydrophilic Surfaces of Cathodically Polarized Amorphous TiO2 Electrodes, J. Electrochem. Soc., 2001, 148, E395 CrossRef CAS.
  58. A. R. Zeradjanin, T. Schilling, S. Seisel, M. Bron and W. Schuhmann, Visualization of chlorine evolution at dimensionally stable anodes by means of scanning electrochemical microscopy, Anal. Chem., 2011, 83, 7645–7650 CrossRef CAS PubMed.
  59. L. R. Bard and A. J. Faulkner, Electrochemical Method: Fundamental and Applications, John Wiley & Sons, Inc., New York, 2nd edn, 2001 Search PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/d0ra09623g

This journal is © The Royal Society of Chemistry 2021