Next Article in Journal
Experimental Evaluation of a Membrane Micro Channel Reactor for Liquid Phase Direct Synthesis of Hydrogen Peroxide in Continuous Flow Using Nafion® Membranes for Safe Utilization of Undiluted Reactants
Previous Article in Journal
Synthesis, Characterization of g-C3N4/SrTiO3 Heterojunctions and Photocatalytic Activity for Organic Pollutants Degradation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Catalyst-Doped Anodic TiO2 Nanotubes: Binder-Free Electrodes for (Photo)Electrochemical Reactions

1
Department of Chemistry and Chemical Engineering, Inha University, 100 Inha-ro, Nam-gu, Incheon 22212, Korea
2
School of Nano and Materials Science and Engineering, Kyungpook National University, 2559 Gyeongsang-daero, Sangju, Gyeongbuk 37224, Korea
*
Authors to whom correspondence should be addressed.
Catalysts 2018, 8(11), 555; https://doi.org/10.3390/catal8110555
Submission received: 31 October 2018 / Revised: 12 November 2018 / Accepted: 13 November 2018 / Published: 17 November 2018

Abstract

:
Nanotubes of the transition metal oxide, TiO2, prepared by electrochemical anodization have been investigated and utilized in many fields because of their specific physical and chemical properties. However, the usage of bare anodic TiO2 nanotubes in (photo)electrochemical reactions is limited by their higher charge transfer resistance and higher bandgaps than those of semiconductor or metal catalysts. In this review, we describe several techniques for doping TiO2 nanotubes with suitable catalysts or active materials to overcome the insulating properties of TiO2 and enhance its charge transfer reaction, and we suggest anodization parameters for the formation of TiO2 nanotubes. We then focus on the (photo)electrochemistry and photocatalysis-related applications of catalyst-doped anodic TiO2 nanotubes grown on Ti foil, including water electrolysis, photocatalysis, and solar cells. We also discuss key examples of the effects of doping and the resulting improvements in the efficiency of doped TiO2 electrodes for the desired (photo)electrochemical reactions.

1. Introduction

Anodization is a facile process that allows the physical and chemical properties of an entire metal surface to be modified at a very low cost, and it has long been recognized as a well-established industrial surface treatment technique [1,2]. In particular, the discovery of self-ordering phenomena in anodic aluminum oxide produced using two-step anodization [3,4,5,6] led to great advances in nanoscience applications based on highly ordered nanoporous Al2O3 membranes and triggered the search for new alternative anodic materials with intrinsic semiconductive properties induced by external electrical or photo stimuli [7].
One such material is anodic titanium oxide. Its morphology, and thus its chemical and physical properties, can be easily tuned by electrochemical anodization [8,9,10,11]. Unlike anodic aluminum oxide (>7.5 eV), anodic titanium oxide has a moderate bandgap (~3.1 eV), can form unique morphologies such as nanotubular, microconical, fishbone-like structures, and can be directly grown on a Ti substrate [9,10,11,12,13,14]. TiO2 itself possesses photocatalytic properties and inherent resistance to corrosion under harsh environmental conditions [8,15,16].
Since the formed oxide adheres to and remains in stable contact with the mother metal material without any binder, high-surface-area anodic titanium oxide has been considered for use as an electrode for (photo)electrochemical reactions in strong acidic or basic conditions [17,18]. In general, electrodes should require a low overpotential for initiating the targeted reaction in order to increase current efficiency. However, oxide-based electrodes containing anodic titanium oxide are disadvantageous in this regard, as their charge transfer resistance is higher than that of metal-based electrodes [19]. Thus, modification of the electrode material with catalysts has been used to enhance charge-transfer at the interface between the electrode and the electrolyte and to lower the overall overpotential of electrodes that are affordable to use in highly corrosive environments or environments requiring long-term stability. For example, electrodes that can effectively split water into hydrogen and oxygen by electrolysis have attracted attention for use in hydrogen generation, a key issue in sustainable energy generation and storage systems [20]. Due to the higher overpotential and sluggishness of the oxygen evolution reaction (OER) in acidic media than that of the hydrogen evolution reaction (HER), the OER is the rate-limiting step for the development of high-efficiency water electrolyzers [20,21,22]. Usually, noble metals are utilized as active catalysts, with Pt being commonly used for the cathode and Ir, Ru, or both being used for the anode. The HER in alkaline media is a more interesting reaction, as non-noble catalysts such as Co, Ni, and Fe work nearly as well as Ir or Ru catalysts [23]. In addition, high-surface-area TiO2 nanotubes containing trace amounts of Ru or Ir catalysts have also been demonstrated to be applicable as OER electrodes in alkaline media [24].
In photo-electrochemistry or photo-catalysis-related applications, the bandgap and band positions are very important, as they determine the energy (or H2) conversion efficiency from solar energy [15,25]. Because the inherent bandgap of TiO2 falls in the UV region, TiO2 nanotubes must be doped with foreign elements (N, S, P, Ru, Ta, Nb, WO3, MoO3, etc) to improve their photocatalytic properties. However, due to the high-aspect-ratio of nanotubes grown on a Ti substrate, uniform and homogenous doping (or decoration) of foreign elements in (or on) the entire surface of the nanotubes is very difficult to achieve.
Very recently, several technical advances for the doping of active materials into high-aspect-ratio TiO2 nanotubes without causing structural damage have been reported by several groups [26,27,28,29,30,31,32]. For example, Ti alloys containing active catalysts were anodized under suitable anodization conditions to produce nanotubular anodic TiO2 containing foreign catalysts [30]. The simplest method involves doping with N or C by annealing TiO2 under a N- or C-providing environment [26,27,28,29]. In addition, anodization can be carried out in a F-based electrolyte containing a negatively charged precursor (for example, RuO4−, which is dissociated from KRuO4) that can be incorporated into the anodic oxide during the anodization, the growth of anodic TiO2 nanotubes occurs and doping with the foreign catalyst occur simultaneously (so-called single-step anodization for doping). Alternatively, immediately after the fabrication of nanotubes, the F-containing electrolyte, which is essential for the formation of nanotubes, can be replaced with an electrolyte containing the negatively charged precursors. Subsequently, a potential higher than that used for anodization is applied to the nanotubes in the precursor solution to dope the negatively charged ions into the oxide (potential shock for doping) [31,32,33].
Although each method has its advantages, both give rise to some issues during the doping process. Specifically, in single-step anodization, some precursors are self-reduced in the electrolyte by F ions. For example, when single-step anodization is carried out using MnO4 in an F-based electrolyte, the Mn is deposited on the counter electrode in metallic form rather than being doped into TiO2 as MnO4 due to the facile reduction of MnO4 to Mn2+ [34]. In the potential shock method, the barrier oxide on the bottom of nanotubes grows significantly thicker with increasing potential shock voltage, blocking current flow. Although higher potentials produce higher doping concentrations in the nanotubes, very high potential shock voltages cannot be used due to the barrier layer growth. Therefore, optimization of the potential shock voltage always involves a trade-off between achieving a high degree of doping and reducing the growth of the barrier oxide. Additionally, although electrolytes with high precursor concentrations result in higher doping concentrations in TiO2, in some cases the target potential cannot be maintained due to the limiting current-compliance of the power supply [34]. Thus, low potential shock voltages and low concentrations of doping precursors are preferred in the potential shock method, meaning that relatively low doping concentrations are achieved. In general, Cl-based ions destroy or break down nanotubular structures during simultaneous single-step anodization or potential shock doping. However, some novel metal precursors inherently produce Cl; for example, H2PtCl6 is dissociated to 2H+ + PtCl62−, which generates Cl ions. In such cases, underpotential shock, in which the potential shock voltage is lower than that used for anodization, has been successfully demonstrated as an alternative doping method. In that case, the Pt ions penetrate only the middle of the oxide layer where the low potential shock voltage has an influence, rather than reaching the interface of the barrier oxide and the Ti metal [33].
The anodization of alloys is another method to produce doped TiO2 nanotubes; however, the alloying process is very elaborate compared to the fabrication of Ti metal alone. Due to limit of availability of Ti-based alloys and their cost, a great deal of work to develop easily reproducible metal alloys at low cost is still required [30]. Additionally, the inevitable thermal growth of additional oxides and crystal phase changes as a function of the temperature during the annealing process should be considered [35].
Recently, a number of review or prospective articles related to anodic TiO2 and its applications have been published. For example, Lee et al. comprehensively reviewed one-dimensional titanium oxides grown by anodization [8]. More specifically, several groups have published reviews of photo-catalysis or electro-catalysis applications based on anodic TiO2 [36,37,38,39,40]. Additionally, bio-related applications of anodic TiO2 materials have also been reviewed recently [41,42].
In this review, we focus on the use of TiO2 nanotubes as (photo)electrochemical binder-free electrodes in applications that require low overpotential or narrow bandgaps to achieve the desired reactions. Various doping processes to adjust the overpotential and bandgap are comprehensively reviewed.

2. Formation Mechanisms of TiO2 Nanotubes

2.1. Reaction Mechanism

The formation of TiO2 nanotubes is governed by two competitive reactions: (1) oxide formation at the oxide/metal interface by oxygen anions (O2 and OH), and (2) the dissolution reaction caused by the F in the electrolyte (Figure 1) [2,8,10,11,16,36].
In oxide formation, O2 and OH anions evolve on the electrolyte/oxide interface via the electrolysis of water. These anions diffuse through the oxide layer to come into contact with the metal interface. Simultaneously, Ti4+ moves through the oxide layer in the opposite direction, and reacts directly with oxygen anions. Eventually, TiO2 is generated as shown in the equation below [8,11,16,43].
Ti + 2H2O → TiO2 + 4H+ + 4e,
In short, the oxide formation rate is governed by the diffusion and reaction rates of the oxygen anions in the oxide layer. The newly formed oxide pushes up the existing oxide and gradually grows toward the metal substrate. This phenomenon is known as “plastic flow” [11].
Meanwhile, in the dissolution reaction, F ions react with TiO2 directly at the electrolyte/oxide interface to produce [TiF6]2, which is easily dissolved into the electrolyte due to its high solubility in water. As the F ions travel roughly twice as fast as the oxygen ions, they can contact and react directly with Ti4+ at the oxide/metal interface. The relevant reaction equations are follows [8,11,16]:
TiO2 + 4H+ + 6F → [TiF6]2 + H2O,
Ti4+ + 6F → [TiF6]2,
Due to the role of F, the honeycomb-like porous structure transforms into an array of tubes with an F-rich layer between the TiO2 layer and the Ti metal. As a result, a hemispherical oxide/metal interfacial structure is formed. The overall growth rate and the length of the nanotubes are determined by the difference between the oxide formation and dissolution reaction rates.
The current–time (I-t) behavior measured during the potentiostatic preparation of TiO2 nanotubes reflects this overall mechanism (Figure 2). In the initial stage, the current decreases rapidly due to the increasing resistance of the rapidly growing oxide layer (I). In the second step (II), the current increases slightly due to the formation of nano-sized pits, which are formed due to attack by the F ions. The local electric fields are strengthened in these pits, thus attracting more anions to participate in the oxide formation and dissolution reactions. As a tubular structure begins to form beneath the pits, the current decreases again (III). When the oxide formation and dissolution reactions reach equilibrium, the current converges (IV). The current–time transient itself cannot confirm the formation of nano-tubular structures, since non-nanotube-forming processes based on the barrier oxide formation, oxide pitting, and the stable growth mechanism show similar current patterns [11].

2.2. Reaction Parameters

TiO2 nanotubes can be obtained in electrolytes containing halides (ClO4, Br, Cl, F), regardless of the electrolyte solvent. The use of F salts is favorable to prepare uniformly oriented TiO2 nanotube arrays [31,44,45]. When Cl or Br are added to the electrolyte, non-uniform bundle-like nanotubular structures are easily generated; this is known as rapid breakdown anodization (RBA) [30,44,45,46]. Generally, 0.05–2% F ion is considered to be appropriate for the formation of TiO2 nanotubes [11,31]. After selecting the F concentration, the environment of the electrolyte is determined depending on the type of solvent. HF is a typical F source that is widely utilized in aqueous conditions. Additionally, it is recommended to adjusting the pH of electrolyte to neutrality; sulfuric acid, phosphoric acid, and acetic acid are typical acids used for this purpose, and NaOH or KOH are generally used as bases. Shifting the pH of the solution into the weakly acidic range results in longer nanotubes [47,48,49]. Feasible anodization voltages range from 5 to 30 V in aqueous conditions [11].
Ethylene glycol and glycerol are the most commonly used organic solvents for TiO2 nanotube growth. The use of ethylene glycol is favored due to the low glycerol solubility of F containing salts such as NH4F, which are widely used as the F source [50,51]. When these solvents are used, the electrolyte is normally weakly basic, demonstrating that basic conditions can also be used to prepare TiO2 nanotubes. However, water is an essential component for the generation of TiO2. For this reason, 1–10% of water is usually added directly to the organic electrolyte [8,11,16,31,43]. TiO2 nanotubes can be prepared over a very wide range of applied voltages in ethylene glycol; they are easily and uniformly grown via potentiostatic methods without any additional control technique in the range 25–100 V [31,52]. The recently reported ramping method is a powerful solution to avoid the breakdown of the oxide and to form nanotubes evenly at voltages above 100 V under organic conditions [52,53].
The pore size of the TiO2 nanotubes is governed by intensity of the applied voltage. Normally, pores with sizes of 10–100 nm can be produced using the typical voltage ranges given above [11,44,54,55]. For instance, a pore size of around 100 nm is easily obtained at 20 V. Similarly, the size of the pores produced in ethylene glycol solution is proportional to the applied voltage, but a voltage of 40–45 V corresponds to a pore size of 100 nm [31,56,57].
Moreover, it has been reported that the pore size can be expanded to 600 nm or more using the ramping technique, similarly to in the anodization of aluminum [52,53].
If the oxide formation rate is faster than the oxide dissolution rate, the length of the TiO2 nanotubes will increase as a function of anodization time. When the rates of these two competitive reactions are in a steady-state equilibrium, the growth of nanotubes is halted, and they maintain their length. In general, TiO2 nanotubes prepared in an organic electrolyte are much longer than those prepared in aqueous solution. The longest reported nanotubes produced under aqueous conditions were of 7.3 μm in length, but 1–3 μm nanotubes are achieved in most cases [11,31,49,58,59]. In contrast, nanotubes tens or hundreds of μm in length can be grown in ethylene glycol, allowing researchers to adjust the length of the nanotubes from a few hundreds of nanometers to micrometers by controlling the anodization time [8,31,56,57]. However, longer anodization time causes the formation of nanograsses, which are produced by nanotubes splitting in the direction of the applied electric field of the F ions. Since the nanograss blocks the movement of ions into the pores, several methods for its removal have been developed [13,59].

3. Doping of TiO2 Nanotubes

As mentioned earlier, TiO2 nanotubes provide an enormous electrochemical reaction surface area due to their high aspect-ratio. For example, TiO2 nanotubes with a pore diameter of 100 nm and the aspect-ratio of 10 provide a surface area that is 120 times higher (0.785 cm2 → 94.2 cm2) than that of a flat TiO2 surface [31]. Nonetheless, suitable catalysts or active materials must be used to overcome the insulating properties of TiO2. In this section, we discuss methods of doping TiO2 nanotubes.

3.1. Electrochemical Doping Methods

3.1.1. Single-Step Anodization

Single-step anodization is a doping method based on the principle of plastic flow. The salt of an anion complex containing the target metal oxide is selected as the precursor. This salt is added directly to the conventional electrolyte to prepare the nanotubes. As shown in Figure 3a, nanotubes will not be produced if the electrolyte is not suitable for the typical anodic conditions used to produce nanotubular structures [17,24,31,47]. In addition, an electrolyte is difficult to be used if the precursor is vulnerable to F ion [34].
The doping mechanism is almost the same as the anodization mechanism. The precursor anion participates in the two main reaction processes of oxide formation and dissolution. If the target dopant is Ru or Ir oxide, which are widely used for water oxidation catalysts, the catalyst doped into the nanotubes can be partially dissolved via the oxygen evolution reaction during prolonged anodization [17,24,31]. As a result, the doping concentration of the catalyst at the growth site will be much higher than that of the sidewalls of the nanotubes (Figure 3b).
Yoo et al. successfully doped RuO2 into TiO2 nanotubes via single-step anodization in an ethylene-glycol-based electrolyte containing 0.25% NH4F and 5% water. RuO2-doped TiO2 nanotubes with a doping concentration of 0.12 at.% were produced in the same anodizing electrolyte with the addition of 0.02 M KRuO4. The RuO2-doped TiO2 nanotubes prepared at 40 V for 22 h were approximately 30% longer than conventional nanotubes prepared without KRuO4 [17]. The doping concentration increased to 1.21 at.% when an aqueous medium was used due to the effect of water. During the synthesis of the nanotubes, the precursor KRuO4 is dissociated into K+ and RuO4, the latter of which takes part in the following relevant reactions:
Ti + RuO4 → TiO2 + RuO2 + e,
Ti + 1/2RuO4 + H2O → TiO2 + 1/2RuO2 + 2H+ + 5/2e,
As shown in Equations (4) and (5), both RuO2 and TiO2 are formed from RuO4 simultaneously. As a result, the wall thickness of the nanotubes increased, as shown in Figure 4. The wall thickness of the nanotubes increases from 5.73 nm to 13.96 nm in the presence of the catalyst precursor of KRuO4. The thickness can be further increased (13.96 nm → 25.74 nm) when electrolyte contains small amount of water, which can be explained by simultaneous formation of TiO2 and RuO2 in Equations (4) and (5). The rutile RuO2 phase determined by SAED analysis provides a clear evidence for doping (Figure 4d).
Furthermore, simultaneous co-doping of TiO2 nanotubes with IrO2 and RuO2 can be achieved by single-step anodization using the electrolyte 1 M H3PO4 + 1 M NaOH + 0.3–0.7 vol.% HF containing KRuO4 and IrOx nanoparticles, which are prepared as an intermediate species for IrO2 doping. Since chloride ions, which damage the NTs, are typically involved in Ir precursors, IrOx nanoparticles are used to generate IrO4 in the electrolyte along with the RuO4 for RuO2 doping [24].
Ir2O3 + 3H2O → 2Ir(OH)3,
Ir(OH)3 → IrO3 + 3H+ + 3e,
IrO3 + H2O → IrO4 + 2H+ + e,
Ti + IrO4 → TiO2 + IrO2 + e,
Ti + 1/2IrO4 → TiO2 + 1/2IrO2 + 2H+ + 5/2e,
Therefore, the overall morphology of co-doped TiO2 is similar to that of TiO2 nanotubes singly doped with RuO2. Under the optimized conditions, the doping concentration of IrO2 in the TiO2 layer of binary-catalyst-doped TiO2 is usually about half that of RuO2, exhibiting polycrystalline after annealing.
Tin oxide is also frequently used as a dopant to control the bandgap and overpotential of TiO2. Unlike the doping precursors discussed above, tin oxide, which can be easily obtained from Na2SnO3, is used at basic pH; basic conditions are not typically used for the anodization of Ti in aqueous conditions. Ma et al. reported that an ethylene glycol-based electrolyte containing both Na2SnO3 and Na2MoO4·4H2O led to the formation of TiO2 nanotubes containing both 0.28% MoO3 (Mo6+) and 0.51 at.% SnO2 (Sn4+) [60].
Very recently, we reported that WO3 can also be doped onto TiO2 nanotubes using a phosphoric-acid-based aqueous electrolyte containing Na2WO4. The doping concentration increased linearly with the concentration of the precursor. The WO3-doped TiO2 nanotubes showed 17% higher electrochromic performance than bare TiO2 nanotubes at the optimum doping concentration of 0.21 at.% [47].

3.1.2. Anodic Potential Shock

The anodic potential shock method was first reported by Jo et al. in 2009 [61]. Originally, it was designed for the preparation of through-hole type TiO2 membranes. The same group later adopted this method to introduce dopants in/on anodic structures through modification of the anodization parameters [32,62,63]. This method is regarded as a two-step anodization, with each step being performed in a different electrolyte. In the first step, the formation of nanotubes takes place. In the second step, an extremely high anodic voltage is applied to the prepared nanotubes for a very short period (Figure 1a); the doping precursor is only present in the electrolyte used during the second step. The precursors used are almost the same as those in the single-step anodization, but the oxide formation and doping reactions are independent. Therefore, the oxide formation follows the anodization mechanism exactly (see Figure 1b), and the doping results from high-field migration under a sudden high potential.
We showed that RuO2 could be doped into TiO2 barrier structure via the potential shock method using KRuO4 as the precursor [63]. The average doping concentration of Ru in TiO2 at a shock voltage of 140 V was over 4 at.%. The two-step anodic doping method resulted in a much higher Ru concentration at the base of the nanotubes than in the walls, similarly to in single-step anodization [31,32]. However, additional barrier oxide was formed at the bottom of the tubes due to the absence of F- ions in the doping electrolyte; its thickness was a function of the applied potential shock voltage (see Figure 5). Using a similar method, Seong et al. succeeded in producing TiO2 nanotubes with 0.7 at.% MnO2 from the precursor KMnO4 in ethylene glycol media; the nanotubes showed enhanced water oxidation performance under alkaline conditions [34].
For Pt doping, Cl-based complexes such as H2PtCl6 are typically available as reaction-grade chemicals. Kim et al. prepared TiO2 nanotubes doped with 3 ppm of PtO for application in the hydrogen evolution reaction. In this case, a shock voltage below 20 V was utilized because higher voltages (over 20 V) caused the collapse of the tubular arrays via attack by the Cl ions. Since the applied potential shock voltage was far lower than the anodization voltage, this method is referred to as the underpotential shock method [33].
It is possible to use the potential shock method with any metal oxide catalyst, as long as the precursor is sufficiently soluble in a suitable solvent. However, the entrances to the pores must remain clear to allow the precursors easy access to the deep sites of the nanotubes. For this reason, TiO2 nanotubes prepared in aqueous solution are much preferred. However, the possibility of using potential shock under other conditions has been suggested, as techniques to remove the nanograss that gradually develops during anodization have been reported [64,65,66].

3.2. Doping via Thermal Treatment

Anodically formed TiO2 nanotube layers are normally amorphous [67,68]; however, when anodization is carried under specific conditions at higher voltages, nanocrystallites are present [35,67,69]. However, for use in many applications, the amorphous tubes must be crystallized by thermal treatment. In general, amorphous nanotubes can be converted to the anatase or rutile phases by thermal treatment at 300–500 °C or 550 °C, respectively. The bandgap energy of a crystallized TiO2 nanotube is 3.2 eV after conversion to anatase, and 3.0 eV after conversion to rutile [68]. For certain applications, such as solar cells and photocatalysis, the bandgap energy of these structures is too wide. To enhance their efficiency in such applications, the bandgap energy must be tuned. The most well-known approach for adjusting the bandgap is to dope metal or non-metal impurities into the TiO2 nanotubes.
Asahi et al. reported that nitrogen-doping of TiO2 enhanced its visible photoresponse. Their report demonstrated that the doped nitrogen, which substituted for the oxygen of TiO2, narrowed the bandgap energy by introducing N2p states just above the valence band of TiO2 [15]. The classic approach to oxygen-substitution-doping of TiO2 nanotubes involves ion implantation [26,27,70,71,72]. This method is most effective at introducing nitrogen into the TiO2 lattice at low-to-medium doping levels (about 1018 ions/cm2) [26,27]. However, this method has several shortcomings: the ion penetration depth is limited to a few micrometers, a relatively high acceleration energy of several MeV is required, and the distribution of the dopant in the TiO2 structure is often inhomogeneous. Moreover, after implantation, amorphization of the TiO2 nanotubes occurs, making a reannealing process to reestablish the crystalline structure necessary. A simpler approach for doping nitrogen into TiO2 is thermal treatment in NH3 [70]. Such thermal treatments are generally performed in a NH3/Ar atmosphere at relatively high temperatures (above 500 °C) [29].
Based on the peak position of nitrogen observed from sputtering TiO2 under nitrogen atmosphere and titanium nitride [29], successful nitrogen doping should result in a nitrogen (N1s) peak located at ~396 eV in XPS analysis [28,29,72]. However, in some reports, this nitrogen peak was observed at ~400 eV or above 400 eV [73,74,75]. These higher peak values correspond to surface adsorption or sensitization of nitrogen on the TiO2 surface as N–C compounds or molecular N2 [76]. Several research groups have claimed to have successfully doped nitrogen into TiO2 via solution-based doping. Other reports even claim that the nitrogen doping concentration in TiO2 is increased by prolonged anodization. However, these results require further confirmation, as they resulted in N peaks located at ~400 eV [73,74,75,76]. Moreover, most of these reports did not show suitable evidence of bandgap engineering via visible photocurrent spectra or visible photocatalytic activity.
Carbon has also been suggested as a dopant to form states near the valence band by the substitution of oxygen [25,77,78]. Furthermore, carbon doping can also be achieved by the thermal treatment of TiO2 in a carbonaceous environment or ashing organic compounds, such as CO or acetylene [79,80,81]. The thermal doping of carbon should be carefully defined to differentiate it from graphitization or the formation of oxy-carbides of the TiO2 nanotubes [82,83]. The thermal treatment of TiO2 nanotubes in an acetylene environmental can also convert them to TiOxCy materials with semimetallic conductivity comparable to that of graphite [83]. Solution-based carbon doping is also frequently attempted. However, these results cannot consider carbon residue from decomposition of the organic electrolyte. The organic solvent in the anodization electrolyte can be decomposed by the applied high potential and remain inside the nanotube wall [67,84].

3.3. Alloy-Based Anodization

The simplest and most straightforward doping approach is the anodization of alloys that contain the dopant material. Using this method, nitrogen-doped TiO2 nanotubes can be obtained by the anodization of TiN substrates [85,86], which can be prepared by arc-melting Ti and TiN powders. A similar approach involves anodizing Ti-transition metal alloys to produce metal- or metal-oxide-doped TiO2 nanotubes [87,88,89]. Density functional theory (DFT) calculations have shown that various transition metals are effective dopants for Ti substitution in TiO2 structures to form intermediate states in the bandgap [90]. The substitution of Ti by such dopants leads to a red-shift in the optical properties of the nanotubes [91,92,93,94], enhancing its electrical conductivity [95,96,97,98]. Metal oxides can also substitute Ti, as in the example of W doping. When tungsten is doped at Ti sites, it forms WO3 in the TiO2 lattice. Each W6+ associates with an extra oxygen instead of Ti4+. The WO3-doped TiO2 has a narrower bandgap as a result of its lower conduction band [99].
When preparing metal- or metal-oxide-doped TiO2 nanotubes by the anodization of a Ti alloy, the dopant materials must not interfere with the anodic reaction. That is, in order to successfully prepare porous or tubular structures, only a very small amount of the dopant materials should be present in the Ti alloy, and the dopant materials should have chemical properties similar to those of Ti. If the chemical properties of the dopant are too different from those of Ti, or if too much dopant is present in the Ti alloy, the resulting nanotube structures may be irregular, or in the worst case, only a compact oxide will be formed (Figure 6).
In many cases, interesting morphology such as two-scaled nanotube structures (two length scales or two distinct tube diameters) are formed during the anodization of certain alloys [100,101,102,103,104]. Such morphologies may be caused by the different oxidation or dissolution kinetics of Ti and the dopant metals, but the phenomena are still not completely understood [105,106,107]. The metal composition of the anodic oxide formed from a Ti alloy depends on the metal ratio in the alloy substrate [108]. In some cases, small amounts of mixed oxides may be present in the anodic oxide structures. So far, complete mixed oxide conversion has only been reported for TiZr alloys [109].
Earlier studies of doped TiO2 nanotubes prepared by the anodization of alloys focused on enhancing ion insertion to improve their electrochromic properties [87,88,89,110,111]. In particular, WO3, MO3, and Nb2O5-doped TiO2 nanotubes formed by the anodization of Ti-W, Ti-Mo, and Ti-Nb alloys showed significantly improved electrochromic efficiency and ion (particularly, H+) insertion properties when even a small amount of dopant (0.2 at.% of WO3) was present [87,88,110,111,112,113]. Moreover, TiO2-Nb2O5 mixed-oxide nanotubes showed high ion intercalation stability and electrochromic activity. These results were due to the widening of the TiO2 lattice by Nb ions, which was observed using high-resolution transmission electron microscopy (HRTEM) analysis and predicted by DFT calculations [88]. The lattice-widened TiO2 nanotubes allow lager guest ions such as Li+ and Na+ to intercalate into the lattice, as well as making the H+ intercalation process faster [88].
Obviously, the anodic oxidation of Ti alloys in F-containing electrolytes allows a wide range of dopants to be effectively doped into TiO2 nanotubes. Moreover, the doped TiO2 nanotubes formed from Ti alloys provide virtually unlimited potential to enhance the chemical and physical properties of the TiO2 nanotubes. Specific examples of such enhancements will be discussed in the applications section.

4. Applications

4.1. Electrochemical Water Electrolysis

The electrochemical oxygen evolution reaction (OER) is a four-electron transfer reaction consisting of four single electron transfer reactions. Therefore, the use of catalysts is essential to reduce the overpotential, which is driven by the multiple electron transfer steps [115,116,117]. Increasing the amount of oxygen evolution at a given overpotential is becoming an important research topic [115,118]. A high reaction surface area is an important factor in producing a large amount of oxygen gas. Moreover, a durable support material is required, because the electrode is continuously exposed to corrosive oxygen radicals [119,120]. Thus, TiO2 nanotubes, which have inherent anti-corrosive properties, are expected to be a good electrode material for this application. Decorating or doping catalysts onto/into high-aspect-ratio TiO2 nanotubes is difficult, as the mouths of nanotubes can become blocked by precipitated catalyst if improper methods are employed. However, the problem of pore loss can be effectively solved by preparing the catalyst-assisted TiO2 via the doping methods described in the previous sections.
TiO2 nanotubes that are singly doped with RuO2 via single-step anodization exhibit overpotentials in the range of 770–1000 mV [17,31,32,121]. The ideal Tafel plot slopes for the first, second, and third electron transfer steps in the OER mechanism have been reported to be 120, 40, and 30 mV/dec, meaning that the slope of Tafel should be lower than these values to achieve an excellent electron transfer rate. The Tafel slopes of nanotubes produced by single-step anodization are in the range of 37–46 mV/dec, indicating that the rate-determining step (RDS) for the OER in the nanotubes is located near the second electron transfer step. Nanotubes doped with RuO2 using the potential shock method exhibit a lower overpotential (650 mV) than those produced via single-step anodization. However, the RDS when using these nanotubes for the OER has been reported to be the second electron transfer step, as they have a Tafel slope of 46 mV/dec [31,122,123,124].
The slow electron transfer of electrodes produced via the potential shock method can be explained in terms of morphological changes; the increased thickness of the barrier oxide at the bottom of the tubes increases the electron transfer resistance [31,54]. However, despite the thick barrier oxide, the overpotential of nanotubes prepared by the potential shock is lower than that of those prepared by single-step anodization. This is due to the surface chemical composition of the catalysts. Ru-O bonding, which effectively diminishes the overpotential, is dominant on the surface of the potential shock nanotubes [125,126,127]. On the other hand, Ru-OH, which increases the overpotential, is dominant on the surface of the single-step anodization nanotubes [31,121].
A complex variety of factors determine the current density in the OER. Increased surface area and an increased doping concentration of Ru in TiO2 lead to higher current density, whereas increased barrier layer thickness reduces the current density (Figure 7). Overall, based on empirical experimental data from RuO2-doped TiO2 nanotubes, we found that the OER current is determined by three major factors according to the following relationship:
  i     e ( η ) S   ×   C R u 2 B 1.5  
  • i = end current density of LSV for OER
  • η = overpotential (0.1 M KOH vs. RHE)
  • CRu = average concentration of Ru determined by TEM EDS
  • S = calculated reaction surface area of electrode
  • B = average thickness of barrier oxide
TiO2 nanotubes doped with MnO2 via potential shock have been used as an improved non-noble metal catalyst in alkaline OER electrodes [34]. Pt, specifically PtO, which is doped into TiO2 nanotubes by underpotential shock, provides enhanced performance in both the OER and HER (hydrogen evolution reaction) [33]. RuO2 and IrO2 co-doping of TiO2 nanotubes shows a synergetic effect in reducing overpotential. Such co-doped nanotubes exhibited an overpotential of 590 mV with the third electron transfer step [128]. This electrode evolved twice the amount of oxygen gas compared to a powder-type electrode material with identical components (Figure 8) [24,128,129].

4.2. Photoanodes

4.2.1. Photocatalysis and Photoelectrochemical Water Splitting

Since the first report of photocatalytic water splitting on TiO2 by Fujishima and Honda in 1972 [130], TiO2 has been considered to be one of the best materials for photocatalysts. It has excellent photooxidative activity for the degradation of organic pollutants, toxins, and bacteria [130,131,132,133,134,135,136]. Two main reactions occur simultaneously in a photocatalytic system: photooxidation from holes on the surface of the valence band edge and photoreduction from the electrons on the surface of conduction bad edge.
Enormous efforts have been dedicated to the study of the physical, chemical, and material properties of TiO2 photocatalysts in order to enhance their activity. Recently, several investigations reported that anodic TiO2 nanotube layers are more promising than nanoparticle layers for improving photocatalytic efficiency due to their well-defined geometry [67,79,137,138,139,140,141,142,143,144,145,146] and the ability to easily incorporate catalysts and dopants [141,147,148,149,150,151].
Among the various metal- and metal-oxide-doped TiO2 nanotubes, WO3-doped TiO2 nanotubes most effectively increase photocatalytic activity [141]. In contrast to Al-doped nanotubes (one of the most efficient additives for inducing carrier recombination [89,141]), mixed-oxide TiO2 nanotubes doped with both WO3 and MoO3 show strongly enhanced photocatalytic activity compared with non-doped tubes. The highly beneficial effect of W and Mo cannot be explained by better charge transport in the tubes but is instead ascribed to modification of the band or surface state distribution of the doped nanotubes [89,112,141].
Photoelectrochemical water splitting is another promising application for anodic TiO2 nanotubes. Like photocatalysis, the photoelectrochemical water splitting reaction is based on the light-induced electron–hole pair creation [140]. H2O can be oxidized by the hole generated on the TiO2 surface, and hydrogen evolution occurs at the counter electrode (such as Pt) by attracting an electron from the TiO2 conduction band. In this context, the electronic properties of TiO2 are very important, because they determine how efficiently electrons can be transferred along the one-directional path. Hence, a wide range of investigations into the electrical and optical properties of TiO2 nanotubes have been performed, as described above.
For photoelectrochemical water splitting, very low amounts of metal or metal oxide dopants such as RuO2 [149], Nb [148,150], and Ta [151] are used. In RuO2-doped TiO2 nanotubes, the RuO2 is believed to accelerate the O2 evolution reaction by catalytic activity [152,153,154]. On the other hand, TiO2 nanotubes doped with small amounts of transition metal (such as Nb or Ta) show enhanced the electrical conductivity due to their narrower bandgaps, meaning that the electrons can travel efficiently through the TiO2 layers to the back contact. The long electron lifetime significantly increases the water splitting efficiency [148,151].

4.2.2. Solar Cells

Another highly promising application of TiO2 nanotubes is energy conversion devices such as dye-sensitized solar cells (DSSCs). In 1991, Grätzel and O’Regan reported the most significant achievement in this field, the first report of fully fabricated solar cell devices (DSSCs) [155]. The solar cells consisted of a nanocrystalline mesoporous TiO2 thin-film electrode, a Ru−bipyridyl complex, and an iodine redox electrolyte, and showed a conversion efficiency of 11% [156,157].
As illustrated in Figure 9, electrons are excited from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO) of the sensitized dye on the TiO2 surface. The excited electrons are injected into the conduction band of the metal oxide on a femto- to picosecond timescale, and the oxidized dye molecules are reduced by the electrolyte redox reaction within nanoseconds. However, the electron transport rate in TiO2 and the diffusion rate of the electrolyte are quite slow (micro- to milliseconds) (Figure 9a). For this reason, the overall cell efficiency is determined by the electron transport rate and electrolyte diffusion rate [158]. The electron transport rate and electrolyte diffusion rate compete with the recombination rate of the electrons. Generally, the electron transport rate in TiO2 nanoparticles is considered to be relatively slow due to surface states, defects, and grain boundaries, which act as electron trapping sites and recombination sites [159,160,161,162,163].
In order to overcome the drawbacks of TiO2 nanoparticles, one-dimensional TiO2 nanostructures such as nanorods, nanowires, and nanotubes have been considered for use as photoanodes for DSSCs. Among these nanostructures, anodically formed TiO2 nanotubes have been considered to be one of the most promising approaches to achieve vertically oriented nanostructures that lead to fast electron pathways [164,165,166,167]. However, the overall conversion efficiency of solar cells based on TiO2 nanotubes is far lower than that of those based on classical nanoparticles. Anodic TiO2 nanotube structures still have considerable room for improvement. Several approaches to enhance the overall conversion efficiency of DSSCs by modifying the nanotube structures have been considered. For example, the charge collection efficiency of nanotubes can be improved by electronic, surface, or geometric modification, which directly affects the amount of dye absorption. In addition, front-side illuminated cell construction obviously enhances light harvesting.
Metal (such as Nb, Ta or Ru) doping enhances the electrical conductivity of TiO2 nanotubes for use in DSSCs [114,168,169]. Doped TiO2 nanotubes have a slower recombination rate than bare TiO2 nanotubes, improving the electron lifetime [114,168,169]. This beneficial effect has been demonstrated using intensity modulated photovoltage spectroscopy (IMVS), photocurrent spectroscopy (IMPS) measurements, and classical electrochemical impedance spectroscopy [114,168,169] (Figure 10). As a result, the overall conversion efficiency of doped TiO2 nanotubes can be increased by 15–35% compared to non-doped TiO2 nanotubes [114,168,169]. Interestingly, these positive effects vanish in nanostructures with high concentrations of the metal dopants.

5. Summary

This paper reviewed the growth of catalyst-doped anodic TiO2 nanotubes and their applications as binder-free electrodes for highly desirable (photo)electrochemical reactions that require a low overvoltage or bandgap. A broad overview of techniques for doping the target materials into anodic TiO2 nanotubes is given.
Concerning the growth of anodic TiO2 nanotubes, we addressed the mechanisms of the reactions that occur during the anodization process and the reaction parameters that influence the structural morphologies of the tubes. The formation of anodic TiO2 nanotubes is guided by oxide formation and dissolution at the interface between the oxide and metal substrate based on a “plastic flow model”. In general, electrolytes containing halide ions are suitable for the preparation of TiO2 nanotubes; in particular, electrolytes containing 0.02–2% F ions are widely used with a constant voltage of 5–30 V in aqueous conditions or 25–100 V in organic conditions.
Several techniques for doping foreign elements into high-aspect-ratio TiO2 nanotubes to improve their catalytic properties have also been reviewed in detail. In single-step anodization, a salt of the anion complex containing the target metal oxide is used as the doping precursor, and is added directly to the electrolyte used for the formation of TiO2 nanotubes under conventional voltage conditions. As the negatively charged precursor is incorporated into the anodic oxide during anodization, the growth of TiO2 nanotubes and doping with the target element occur simultaneously. In the anodic potential shock method, the formation of the TiO2 nanotubes and doping with the target element are performed independently in different electrolytes. After the TiO2 nanotubes are fabricated, the electrolyte is replaced with an electrolyte containing the anion complex precursor and a potential higher than that of first anodization potential is applied for a short time, which leads to the doping of the negatively charged materials into the anodic oxide.
In addition to electrochemical doping methods, simple and straightforward doping approaches, such as the thermal annealing of TiO2 nanotubes under a suitable environment or the anodization of an alloy containing the dopant materials can be employed.
Catalyst-doped anodic TiO2 nanotubes lead to low overpotentials, which are favorable for initiating reactions. For this reason, they are considered to be promising electrodes for electrochemical applications. The overpotential for electrochemical water electrolysis can be reduced to 650 mV via doping of TiO2 nanotubes. Due to the morphological changes and surface composition of the catalyst on the electrode surface, it is difficult to further reduce the overvoltage for OER.
In addition, metal or metal oxide (Nb, Ta, Ru, WO3 or MoO3)-doped TiO2 nanotubes show superior photocatalytic and photoelectrochemical water splitting performance compared to non-doped TiO2 nanotubes due to bandgap engineering and their surface state distribution. Nb, Ta, or Ru-doped TiO2 nanotubes show higher overall conversion efficiency in DSSCs as a result of their slower recombination rate and enhanced electron lifetime.

Author Contributions

Supervision: K.L. and J.C.; Writing—original draft preparation: H.Y., K.L., M.K., Y.-T.K. and J.C.: Writing—review and editing, H.Y., K.L., M.K., Y.-T.K., and J.C.

Funding

This research was supported in part by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (NRF-2018R1A6A1A03024962), and in partly funded through the Human Resources Development program (Grant No. 20174030201500) of the Korea Institute of Energy Technology Evaluation and Planning (KETEP) by the Ministry of Trade, Industry and Energy of the Korean government.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Keller, F.; Hunter, M.; Robinson, D.L. Structural features of oxide coatings on aluminum. J. Electrochem. Soc. 1953, 100, 411–419. [Google Scholar] [CrossRef]
  2. Choi, J. Fabrication of monodomain porous alumina using nano-imprint lithography and its applications. Ph. D. Thesis, Martin-Luther-Universität, Halle-Wittenberg, Germany, 2004. [Google Scholar]
  3. Masuda, H.; Fukuda, K. Ordered metal nanohole arrays made by a two-step replication of honeycomb structures of anodic alumina. Science 1995, 268, 1466–1468. [Google Scholar] [CrossRef] [PubMed]
  4. Masuda, H.; Yamada, H.; Satoh, M.; Asoh, H.; Nakao, M.; Tamamura, T. Highly ordered nanochannel-array architecture in anodic alumina. Appl. Phys. Lett. 1997, 71, 2770–2772. [Google Scholar] [CrossRef]
  5. Jessensky, O.; Müller, F.; Gösele, U. Self-organized formation of hexagonal pore arrays in anodic alumina. Appl. Phys. Lett. 1998, 72, 1173–1175. [Google Scholar] [CrossRef]
  6. Nakajima, D.; Kikuchi, T.; Natsui, S.; Suzuki, R.O. Highly ordered anodic alumina nanofibers fabricated via two distinct anodizing processes. ECS Electrochem. Lett. 2015, 4, H14–H17. [Google Scholar] [CrossRef]
  7. Roy, P.; Berger, S.; Schmuki, P. TiO2 nanotubes: Synthesis and applications. Angew. Chem. Int. Ed. 2011, 50, 2904–2939. [Google Scholar] [CrossRef] [PubMed]
  8. Lee, K.; Mazare, A.; Schmiki, P. One-dimensional titanium dioxide nanomaterials: Nanotubes. Chem. Rev. 2014, 114, 9385–9454. [Google Scholar] [CrossRef] [PubMed]
  9. Ge, M.-Z.; Cao, C.-Y.; Huang, J.-Y.; Li, S.-H.; Zhang, S.-N.; Deng, S.; Li, Q.-S.; Zhang, K.-Q.; Lai, Y.-K. Synthesis, modification, and photo/photoelectrocatalytic degradation applications of TiO2 nanotube arrays: A review. Nanotechnol. Rev. 2013, 2, 27–45. [Google Scholar] [CrossRef]
  10. Kowalski, D.; Kim, D.; Schmuki, P. TiO2 nanotubes, nanochannels and mesosponge: Self-organized formation and applications. Nano Today 2013, 8, 235–264. [Google Scholar] [CrossRef]
  11. Lee, K. Anodic Growth of Porous Metal Oxides and Their Applications. Ph.D. Thesis, Friedrich-Alexander-Universität, Erangen-Nürnberg, Germany, 2013. [Google Scholar]
  12. Park, J.; Lee, G.; Choi, J. Key anodization factors for determining the formation of TiO2 microcones vs nanotubes. J. Electrochem. Soc. 2017, 164, D640–D644. [Google Scholar] [CrossRef]
  13. Lim, J.H.; Choi, J. Titanium oxide nanowires originated from anodically-grown nanotubes: The bamboo-splitting-model. Small 2007, 3, 1504–1507. [Google Scholar] [CrossRef] [PubMed]
  14. Riboni, F.; Nguyen, N.T.; So, S.; Schmuki, P. Aligned metal oxide nanotube arrays: Key-aspects of anodic TiO2 nanotube formation and properties. Nanoscale Horiz. 2016, 1, 445–466. [Google Scholar] [CrossRef]
  15. 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] [PubMed]
  16. Schneider, J.; Matuoka, M.; Takeuchi, M.; Zhang, J.; Horiuci, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 photocatalysis: Mechanisms and materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef] [PubMed]
  17. Yoo, H.; Choi, Y.-W.; Choi, J. Ruthenium oxide-doped TiO2 nanotubes by single-step anodization for water-oxidation applications. ChemCatChem 2015, 7, 643–647. [Google Scholar] [CrossRef]
  18. Gong, J.; Lai, Y.; Lin, C. Electrochemically multi-anodized TiO2 nanotube arrays for enhancing hydrogen generation by photoelectrocatalytic water splitting. Electrochim. Acta 2010, 55, 4776–4782. [Google Scholar] [CrossRef]
  19. Trasatti, S. Electrocatalysis: Understanding the success of DSA®. Electrochim. Acta 2000, 45, 2377–2385. [Google Scholar] [CrossRef]
  20. Anantharaj, S.; Ede, S.R.; Sakthikumar, K.; Karkhick, K.; Mishra, S.; Kundu, S. Recent trends and perspectives in electrochemical water splitting with an emphasis on sulfide, selenide, and phosphide catalysts of Fe, Co, and Ni: A review. ACS Catal. 2016, 6, 8069–8097. [Google Scholar] [CrossRef]
  21. Matsumoto, Y.; Sato, E. Electrocatalytic properties of transition metal oxides for oxygen evolution reaction. Mater. Chem. Phys. 1986, 15, 397–426. [Google Scholar] [CrossRef]
  22. Bockris, J.O.; Otagawa, T. Mechanism of oxygen evolution on perovskites. J. Phys. Chem. 1983, 87, 2960–2971. [Google Scholar] [CrossRef]
  23. Rungtaweevoranit, B.; Zhao, Y.; Choi, K.M.; Yaghi, O.M. Cooperative effects at the interface of nanocrystalline metal-organic frameworks. Nano Res. 2016, 9, 47–58. [Google Scholar] [CrossRef]
  24. Yoo, H.; Oh, K.; Lee, Y.R.; Row, K.H.; Lee, G.; Choi, J. Simultaneous co-doping of RuO2 and IrO2 into anodic TiO2 nanotubes: A binary catalysts for electrochemical water splitting. Int. J. Hydrogen Energy 2017, 42, 6657–6664. [Google Scholar] [CrossRef]
  25. Murphy, A.B. Does carbon doping of TiO2 allow water splitting in visible light? Comments on “Nanotube enhanced photoresponse of carbon modified (CM)-n-TiO2 for efficient water splitting”. Sol. Energy Mater. Solar Cells 2008, 92, 363–367. [Google Scholar] [CrossRef]
  26. Ghicov, A.; Macak, J.M.; Tsuchiya, H.; Kunze, J.; Haeublein, V.; Frey, L.; Schmuki, P. Ion implantation and annealing for an efficient N-doping of TiO2 nanotubes. Nano Lett. 2006, 6, 1080–1082. [Google Scholar] [CrossRef]
  27. Ghicov, A.; Macak, J.M.; Tsuchiya, H.; Kunze, J.; Haeublein, V.; Kleber, S.; Schmuki, P. TiO2 nanotube layers: Dose effects during nitrogen doping by ion implantation. Chem. Phys. Lett. 2006, 419, 426–429. [Google Scholar] [CrossRef]
  28. Macak, J.M.; Ghicov, A.; Hahn, R.; Tsuchiya, H.; Schmuki, P. Photoelectrochemical properties of N-doped self-organized titania nanotube layers with different thicknesses. J. Mater. Res. 2006, 21, 2824–2828. [Google Scholar] [CrossRef] [Green Version]
  29. Vitiello, R.P.; Macak, J.M.; Ghicov, A.; Tsuchiya, H.; Dick, L.F.P.; Schmuki, P. N-Doping of anodic TiO2 nanotubes using heat treatment in ammonia. Electrochem. Commun. 2006, 8, 544–548. [Google Scholar] [CrossRef]
  30. Jha, H.; Hahn, R.; Schmuki, P. Ultrafast oxide nanotube formation on TiNb, TiZr and TiTa alloys by rapid breakdown anodization. Electrochim. Acta 2010, 55, 8883–8887. [Google Scholar] [CrossRef]
  31. Yoo, H. TiO2-based electrodes for electrochemical energy conversion and storage. Ph.D. Thesis, Inha University, Incheon, South Korea, August 2018. [Google Scholar]
  32. Gim, Y.; Seong, M.; Choi, Y.-W.; Choi, J. RuO2-doping into high-aspect-ratio anodic TiO2 nanotubes by electrochemical potential shock for water oxidation. Electrochem. Commun. 2015, 52, 37–40. [Google Scholar] [CrossRef]
  33. Kim, S.; Yoo, H.; Rhee, O.; Choi, J. Doping of Pt into anodic TiO2 nanotubes for water oxidation: Underpotential shock method in Cl solution. J. Phys. Chem. C 2015, 119, 21497–21503. [Google Scholar] [CrossRef]
  34. Seong, M.; Kim, S.; Yoo, H.; Choi, J. Doping of anodic nanotubular TiO2 electrodes with MnO2 for use as catalysts in water oxidation. Catal. Today 2016, 260, 135–139. [Google Scholar] [CrossRef]
  35. Albu, S.P.; Tsuchiya, H.; Fujimoto, S.; Schmuki, P. TiO2 nanotubes—Annealing effects on detailed morphology and structure. Eur. J. Inorg. Chem. 2010, 2010, 4351–4356. [Google Scholar] [CrossRef]
  36. Choi, J.; Wehrspohn, R.B.; Lee, J.; Gösele, U. Anodization of nanoimprinted titanium: A comparison with formation of porous alumina. Electrochim. Acta 2004, 49, 2645–2652. [Google Scholar] [CrossRef]
  37. Paramasivam, I.; Jha, H.; Liu, N.; Schmuki, P. A review of photocatalysis using self-organized TiO2 nanotubes and other ordered oxide nanostructures. Small 2012, 8, 3073–3103. [Google Scholar] [CrossRef] [PubMed]
  38. Smith, Y.R.; Ray, R.S.; Carlson, K.; Sarma, B.; Misra, M. Self-ordered titanium dioxide nanotube arrays: Anodic synthesis and their photo/electro-catalytic applications. Materials 2013, 6, 2892–2957. [Google Scholar] [CrossRef] [PubMed]
  39. Altomare, M.; Nguyen, N.T.; Schmuki, P. Templated dewetting: Designing entirely self-organized platforms for photocatalysis. Chem. Sci. 2016, 7, 6865–6886. [Google Scholar] [CrossRef] [PubMed]
  40. Po, C.C.; Chien, C.C.; Shih, H.C. A review on production, characterization, and photocatalytic applications of TiO2 nanoparticles and nanotubes. Curr. Nanosci. 2017, 13, 373–393. [Google Scholar] [CrossRef]
  41. Fu, Y.; Mo, A. A review on the electrochemically self-organized titania nanotube arrays: Synthesis, modifications, and biomedical applications. Nanoscale Res. Lett. 2018, 13, 187. [Google Scholar] [CrossRef] [PubMed]
  42. Awad, N.K.; Edwards, S.L.; Morsi, Y.S. A review of TiO2 NTs on Ti metal: Electrochemical synthesis, functionalization and potential use as bone implants. Mater. Sci. Eng. C 2017, 76, 1401–1412. [Google Scholar] [CrossRef] [PubMed]
  43. Roy, P.; Kim, D.; Lee, K.; Spiecker, E.; Schmuki, P. TiO2 nanotubes and their application in dye-sensitized solar cells. Nanoscale 2010, 2, 45–59. [Google Scholar] [CrossRef] [PubMed]
  44. Kim, S.; Seong, M.; Choi, J. Rapid breakdown anodization for the preparation of titania nanotubes in halogen-free acid. J. Electrochem. Soc. 2015, 162, C205–C208. [Google Scholar] [CrossRef]
  45. Hahn, R.; Macák, J.M.; Schmuki, P. Rapid anodic growth of TiO2 and WO3 nanotubes in fluoride free electrolytes. Electrochem. Commun. 2007, 9, 947–957. [Google Scholar] [CrossRef]
  46. Nguyen, Q.A.; Bhargava, Y.V.; Devine, T.M. Titania nanotube formation in chloride and bromide containing electrolytes. Electrochem. Commun. 2008, 10, 471–475. [Google Scholar] [CrossRef]
  47. Yoo, H.; Oh, K.; Nah, Y.-C.; Choi, J.; Lee, K. Single-step anodization for formation of WO3-doped TiO2 nanotubes toward enhanced electrochromic performance. ChemElectroChem 2018, 5, 3379–3382. [Google Scholar] [CrossRef]
  48. Macák, J.M.; Tsuchiyam, H.; Schmuki, P. High-aspect-ratio TiO2 nanotubes by anodization of titanium. Angew. Chem. Int. Ed. 2005, 44, 2100–2102. [Google Scholar] [CrossRef] [PubMed]
  49. Lee, B.G.; Nam, S.-C.; Choi, J. Anodic TiO2 nanotubes as anode electrode in Li-air and Li-ion batteries. Curr. Appl. Phys. 2012, 12, 1580–1585. [Google Scholar] [CrossRef]
  50. Paulose, M.; Shankar, K.; Yoriya, S.; Prakasam, H.E.; Varghese, O.K.; Mor, G.K.; Latempa, T.A.; Fitzgerald, A.; Grimes, C.A. Anodic growth of highly ordered TiO2 nanotube arrays to 134 µm in length. J. Phys. Chem. B 2006, 110, 16179–16184. [Google Scholar] [CrossRef] [PubMed]
  51. Shankar, K.; Mor, G.; Fitzgerald, K.A.; Grimes, C.A. Cation effect on the electrochemical formation of very high aspect ratio TiO2 nanotube arrays in formamide−water mixtures. J. Phys. Chem. C 2007, 111, 21–26. [Google Scholar] [CrossRef]
  52. Yin, H.; Liu, H.; Shen, W.Z. The large diameter and fast growth of self-organized TiO2 nanotube arrays achieved via electrochemical anodization. Nanotechnol. 2010, 21, 035601. [Google Scholar] [CrossRef] [PubMed]
  53. Elzarka, A.; Liu, N.; Hwang, I.; Kamal, M.; Schmuki, P. Large-diameter TiO2 nanotubes enable wall engineering with conformal hierarchical decoration and blocking layers for enhanced efficiency in dye-sensitized solar cells (DSSC). Chem. Eur. J. 2017, 23, 12995–12999. [Google Scholar] [CrossRef] [PubMed]
  54. Lee, K.; Schmuki, P. Bottom sealing and photoelectrochemical properties of different types of anodic TiO2 nanotubes. Electrhcim. Acta 2013, 100, 229–235. [Google Scholar] [CrossRef]
  55. Chen, X.; Schriver, M.; Suen, T.; Mao, S.S. Fabrication of 10 nm diameter TiO2 nanotube arrays by titanium anodization. Thin Solid Films 2007, 515, 8511–8514. [Google Scholar] [CrossRef]
  56. Hoseinzadeh, T.; Ghorannevis, Z.; Ghoranneviss, M.; Sari, A.H.; Salem, M.K. Effects of various applied voltages on physical properties of TiO2 nanotubes by anodization method. J. Theor. Appl. Phys. 2017, 11, 243–248. [Google Scholar] [CrossRef]
  57. Atyaoui, A.; Cachet, H.; Sutter, E.M.M.; Bousselmi, L. Effect of the anodization voltage on the dimensions and photoactivity of titania nanotubes arrays. Surf. Interface Anal. 2013, 45, 1751–1759. [Google Scholar] [CrossRef]
  58. Cha, G.; Lee, H.J.; Choi, J. Preparation of binder-free thin film Li4Ti5O12 anode with an adjustable thickness through anodic TiO2 nanotubes. Curr. Appl. Phys. 2013, 13, 1788–1795. [Google Scholar] [CrossRef]
  59. Kim, D.; Ghicov, A.; Schmuki, P. TiO2 Nanotube arrays: Elimination of disordered top layers (“nanograss”) for improved photoconversion efficiency in dye-sensitized solar cells. Electrochem. Commun. 2008, 10, 1835–1838. [Google Scholar] [CrossRef]
  60. Ma, X.; Sun, Z.; Hu, X. Synthesis of tin and molybdenum co-doped TiO2 nanotube arrays for the photoelectrocatalytic oxidation of phenol in aqueous solution. Mater. Sci. Semicond. Proc. 2018, 85, 150–159. [Google Scholar] [CrossRef]
  61. Jo, Y.; Jung, I.; Lee, I.; Choi, J.; Tak, Y. Fabrication of through-hole TiO2 nanotubes by potential shock. Electrochem. Comm. 2010, 12, 616–619. [Google Scholar] [CrossRef]
  62. Shin, S.; Kim, K.; Choi, J. Fabrication of ruthenium-doped TiO2 electrodes by one-step anodization for electrolysis applications. Electrochem. Commun. 2013, 36, 88–91. [Google Scholar] [CrossRef]
  63. Shin, S.; Choi, Y.-W.; Choi, J. Water splitting by dimensionally stable anode prepared through micro-arc oxidation. Mater. Lett. 2013, 105, 117–119. [Google Scholar] [CrossRef]
  64. Wei, W.; Oltean, G.; Tai, C.-W.; Edström, K.; Björefors, F.; Nyholm, L. Hogh energy and power density TiO2 nanotube electrodes for 3D Li-ion microbatteries. J. Mater. Chem. A 2013, 1, 8160–8170. [Google Scholar] [CrossRef]
  65. Sitler, S.J.; Raja, K.S.; Karmiol, Z.; Chidambaram, D. Photoelectrochemical characterization of dual-layered anodic TiO2 nanotubes with honeycomb morphology. J. Phys. D Appl. Phys. 2016, 50, 035502. [Google Scholar] [CrossRef]
  66. Wang, X.; Zhang, S.; Sun, L. A Two-step anodization to grow high-aspect-ratio TiO2 nanotubes. Thin Solid Films 2011, 519, 4694–4698. [Google Scholar] [CrossRef]
  67. Albu, S.P.; Ghicov, A.; Aldabergenova, S.; Drechsel, P.; LeClere, D.; Thompson, G.E.; Macak, J.M.; Schmuki, P. Formation of double-walled TiO2 nanotubes and robust anatase membranes. Adv. Mater. 2008, 20, 4135–4139. [Google Scholar] [CrossRef]
  68. Beranek, R.; Tsuchiya, H.; Sugishima, T.; Macak, J.M.; Taveira, L.; Fujimoto, S.; Kisch, H.; Schmuki, P. Enhancement and limits of the photoelectrochemical response from anodic TiO2 nanotubes. Appl. Phys. Lett. 2005, 87, 243114. [Google Scholar] [CrossRef]
  69. Beranek, R.; Hildebrand, H.; Schmuki, P. Self-organized porous titanium oxide prepared in H2SO4/HF electrolytes. Electrochem. Sol. State Lett. 2003, 6, B12–B14. [Google Scholar] [CrossRef]
  70. Wang, H.; Lewis, J.P. Second-generation photocatalytic materials: Anion-doped TiO2. J. Phys. Condens. Matter. 2006, 18, 421–434. [Google Scholar] [CrossRef]
  71. Irie, H.; Watanabe, Y.; Hashimoto, K. Nitrogen-concentration dependence on photocatalytic activity of TiO2−xNx powders. J. Phys. Chem. B 2003, 107, 5483–5486. [Google Scholar] [CrossRef]
  72. Zhao, Z.; Liu, Q. Mechanism of higher photocatalytic activity of anatase TiO2 doped with nitrogen under visible-light irradiation from density functional theory calculation. J. Phys. D Appl. Phys. 2008, 41, 025105. [Google Scholar] [CrossRef]
  73. Qiu, X.; Burda, C. Chemically synthesized nitrogen-doped metal oxide nanoparticles. Chem. Phys. Lett. 2007, 339, 1–10. [Google Scholar] [CrossRef]
  74. Prokes, S.M.; Gole, J.L.; Chen, X.; Burda, C.; Carlos, W.E. Defect-related optical behavior in surface modified TiO2 nanostructures. Adv. Funct. Mater. 2005, 15, 161–167. [Google Scholar] [CrossRef]
  75. Liu, S.; Yang, L.; Xu, S.; Luo, S.; Cai, Q. Photocatalytic activities of C–N-doped TiO2 nanotube array/carbon nanorod composite. Electrochem. Commun. 2009, 11, 1748–1751. [Google Scholar] [CrossRef]
  76. Beranek, R.; Macak, J.M.; Gartner, M.; Meyer, K.; Schmuki, P. Enhanced visible light photocurrent generation at surface-modified TiO2 nanotubes. Electrochim. Acta 2009, 54, 2640–2646. [Google Scholar] [CrossRef]
  77. Mazare, A.; Paramasivam, I.; Schmidt-Stein, F.; Lee, K.; Demetrescu, I.; Schmuki, P. Flame annealing effects on self-organized TiO2 nanotubes. Electrochim. Acta 2012, 66, 12–21. [Google Scholar] [CrossRef]
  78. Yang, K.; Dai, Y.; Huang, B.; Whangbo, M.-H. Density functional characterization of the visible-light absorption in substitutional C-anion- and C-cation-doped TiO2. J. Phys. Chem. C 2009, 1, 2624–2629. [Google Scholar] [CrossRef]
  79. Park, J.H.; Kim, S.; Bard, A.J. Novel carbon-doped TiO2 nanotube arrays with high aspect ratios for efficient solar water splitting. Nano Lett. 2006, 6, 24–28. [Google Scholar] [CrossRef] [PubMed]
  80. Hahn, R.; Ghicov, A.; Salonen, J.; Lehto, V.-P.; Schmuki, P. Carbon doping of self-organized TiO2 nanotube layers by thermal acetylene treatment. Nanotechnology 2007, 18, 105604. [Google Scholar] [CrossRef]
  81. Barborini, E.; Conti, A.M.; Kholmanov, I.; Piseri, P.; Podesta, A.; Milani, P.; Cepek, C.; Sakho, O.; Macovez, R.; Sancrotti, M. Nanostructured TiO2 films with 2 eV optical gap. Adv. Mater. 2005, 17, 1842–1846. [Google Scholar] [CrossRef]
  82. Yang, L.; Luo, S.; Liu, S.; Cai, Q. Graphitized carbon nanotubes formed in TiO2 nanotube arrays: A novel functional material with tube-in-tube nanostructure. J. Phys. Chem. C 2008, 112, 8939–8943. [Google Scholar] [CrossRef]
  83. Hahn, R.; Schmidt-Stein, F.; Salonen, J.; Thiemann, S.; Song, Y.Y.; Kunze, J.; Lehto, V.-P.; Schmuki, P. Semimetallic TiO2 nanotubes. Angew. Chem. Int. Ed. 2009, 48, 7236–7239. [Google Scholar] [CrossRef] [PubMed]
  84. Song, Y.-Y.; Roy, P.; Paramasivam, I.; Schmuki, P. Voltage-induced payload release and wettability control on TiO2 and TiO2 nanotubes. Angew. Chem. Int. Ed. 2010, 49, 351–354. [Google Scholar] [CrossRef] [PubMed]
  85. Aruna, S.T.; Tirosh, S.; Zaban, A. Nanosize rutile titania particle synthesis via a hydrothermal method without mineralizers. J. Mater. Chem. 2000, 10, 2388–2391. [Google Scholar] [CrossRef]
  86. Kim, D.; Tsuchiya, H.; Fujimoto, S.; Schmidt-Stein, F.; Schmuki, P. Nitrogen-doped TiO2 mesosponge layers formed by anodization of nitrogen-containing Ti alloys. J. Solid State Electrochem. 2012, 16, 89–92. [Google Scholar] [CrossRef]
  87. Nah, Y.-C.; Ghicov, A.; Kim, D.; Berger, S.; Schmuki, P. TiO2−WO3 Composite nanotubes by alloy anodization: Growth and enhanced electrochromic properties. J. Am. Chem. Soc. 2008, 130, 16154–16155. [Google Scholar] [CrossRef] [PubMed]
  88. Ghicov, A.; Yamamoto, M.; Schmuki, P. Lattice Widening in niobium-doped TiO2 nanotubes: Efficient ion intercalation and swift electrochromic contrast. Angew. Chem. Int. Ed. 2008, 47, 7934–7937. [Google Scholar] [CrossRef] [PubMed]
  89. Agarwal, P.; Paramasivam, I.; Shrestha, N.K.; Schmuki, P. MoO3 in Self-organized TiO2 nanotubes for enhanced photocatalytic activity. Chem. Asian J. 2010, 5, 66–69. [Google Scholar] [CrossRef] [PubMed]
  90. Czoska, A.; Livraghi, S.; Chiesa, M.; Giamello, E.; Agnoli, S.; Granozzi, S.; Finazzi, E.; Valentin, C.D.; Pacchioni, G. The nature of defects in fluorine-doped TiO2. J. Phys. Chem. C 2008, 112, 8951–8956. [Google Scholar] [CrossRef]
  91. Osorio-Guillen, J.; Lany, S.; Zunger, A. Atomic control of conductivity versus ferromagnetism in wide-gap oxides via selective doping: V, Nb, Ta in anatase TiO2. Phys. Rev. Lett. 2008, 100, 036601. [Google Scholar] [CrossRef] [PubMed]
  92. Wang, Y.; Doren, D. Electronic structures of V-doped anatase TiO2. Solid State Commun. 2005, 136, 142–146. [Google Scholar] [CrossRef]
  93. Valentin, C.D.; Pacchioni, G.; Onishi, H.; Kudo, A. Cr/Sb co-doped TiO2 from first principles calculations. Chem. Phys. Lett. 2009, 469, 166–171. [Google Scholar] [CrossRef]
  94. Shao, G.J. Red shift in manganese- and iron-doped TiO2: A DFT+U analysis. Phys. Chem. C 2009, 113, 6800–6808. [Google Scholar] [CrossRef]
  95. Hotsenpiller, P.A.M.; Bolt, J.D.; Farneth, W.E.; Lowekamp, J.B.; Rohrer, G.S. Orientation dependence of photochemical reactions on TiO2 surfaces. J. Phys. Chem. B 1998, 102, 3216–3226. [Google Scholar] [CrossRef]
  96. Casarin, M.; Maccato, C.; Vittadini, A. Electronic structure of Nb impurities in and on TiO2. Phys. Chem. Chem. Phys. 1999, 1, 3793–3799. [Google Scholar] [CrossRef]
  97. Atashbar, M.Z.; Sun, H.T.; Gong, B.; Wlodarski, W.; Lamb, R. XPS study of Nb-doped oxygen sensing TiO2 thin films prepared by sol-gel method. Thin Solid Films 1998, 326, 238–244. [Google Scholar] [CrossRef]
  98. Stodolny, M.; Laniecki, M. Synthesis and characterization of mesoporous Ta2O5–TiO2 photocatalysts for water splitting. Catal. Today 2009, 142, 314–319. [Google Scholar] [CrossRef]
  99. Fernandez-Garcia, M.; Martinez-Arias, A.; Fuerte, A.; Conesa, J.C. Nanostructured Ti−W mixed-metal oxides:  Structural and electronic properties. J. Phys. Chem. B 2005, 109, 6075–6083. [Google Scholar] [CrossRef] [PubMed]
  100. Tsuchiya, H.; Macak, J.M.; Ghlcov, A.; Schmuki, P. Self-organization of anodic nanotubes on two size scales. Small 2006, 2, 888–891. [Google Scholar] [CrossRef] [PubMed]
  101. Tsuchiya, H.; Akaki, T.; Terada, D.; Tsuji, N.; Minamino, Y.; Schmuki, P.; Fujimoto, S. Metallurgical aspects on the formation of self-organized anodic oxide nanotube layers. Electrochim. Acta 2009, 54, 5155–5162. [Google Scholar] [CrossRef]
  102. Macak, J.M.; Tsuchiya, H.; Ghicov, A.; Yasuda, K.; Hahn, R.; Bauer, S.; Schmuki, P. TiO2 nanotubes: Self-organized electrochemical formation, properties and applications. Curr. Opin. Solid State Mater. Sci. 2007, 11, 3–18. [Google Scholar] [CrossRef]
  103. Tsuchiya, H.; Akaki, T.; Nakata, J.; Terada, D.; Tsuji, N.; Koizumi, Y.; Minamino, Y.; Schmuki, P.; Fujimoto, S. Anodic oxide nanotube layers on Ti–Ta alloys: Substrate composition, microstructure and self-organization on two-size scales. Corros. Sci. 2009, 51, 1528–1533. [Google Scholar] [CrossRef]
  104. Feng, X.J.; Macak, J.M.; Albu, S.P.; Schmuki, P. Electrochemical formation of self-organized anodic nanotube coating on Ti–28Zr–8Nb biomedical alloy surface. Acta Biomater. 2008, 4, 318–323. [Google Scholar] [CrossRef] [PubMed]
  105. Yasuda, K.; Schmuki, P. Control of morphology and composition of self-organized zirconium titanate nanotubes formed in (NH4)2SO4/NH4F electrolytes. Electrochim. Acta 2007, 52, 4053–4061. [Google Scholar] [CrossRef]
  106. Yasuda, K.; Macak, J.M.; Berger, S.; Ghicov, A.; Schmuki, P. Mechanistic aspects of the self-organization process for oxide nanotube formation on valve metals. J. Electrochem. Soc. 2007, 154, C472–C478. [Google Scholar] [CrossRef]
  107. Yasuda, K.; Schmuki, P. Electrochemical formation of self-organized zirconium titanate nanotube multilayers. Electrochem. Commun. 2007, 9, 615–619. [Google Scholar] [CrossRef]
  108. Kamkin, A.N.; Fishgoit, L.A.; Davydov, A.D. Composition and structure of anodic oxide films on titanium–aluminum alloys by fast electron reflection diffraction, rutherford backscattering, and secondary neutral particle mass spectrometry. Russ. J. Electrochem. 2003, 39, 665–670. [Google Scholar] [CrossRef]
  109. Yasuda, K.; Schmuki, P. Formation of self-organized zirconium titanate nanotube layers by alloy anodization. Adv. Mater. 2007, 19, 1757–1760. [Google Scholar] [CrossRef]
  110. Nah, Y.C.; Ghicov, A.; Kim, D.; Schmuki, P. Enhanced electrochromic properties of self-organized nanoporous WO3. Electrochem. Commun. 2008, 10, 1777–1780. [Google Scholar] [CrossRef]
  111. Shrestha, N.K.; Nah, Y.C.; Tsuchiya, H.; Schmuki, P. Self-organized nano-tubes of TiO2-MoO3 with enhanced electrochromic properties. Chem. Commun. 2009, 15, 2008–2010. [Google Scholar] [CrossRef] [PubMed]
  112. Benoit, A.; Paramasivam, I.; Nah, Y.C.; Roy, P.; Schmuki, P. Decoration of TiO2 nanotube layers with WO3 nanocrystals for high-electrochromic activity. Electrochem. Commun. 2009, 11, 728–732. [Google Scholar] [CrossRef]
  113. Yang, M.; Shrestha, N.K.; Schmuki, P. Thick porous tungsten trioxide films by anodization of tungsten in fluoride containing phosphoric acid electrolyte. Electrochem. Commun. 2009, 11, 1908–1911. [Google Scholar] [CrossRef]
  114. So, S.; Lee, K.; Schmuki, P. Ru-doped TiO2 nanotubes: Improved performance in dye-sensitized solar cells. Phys. Status Solidi RRL 2012, 6, 169–171. [Google Scholar] [CrossRef]
  115. Doyle, R.L.; Lyons, M.E.G. The Oxygen Evolution Reaction: Mechanistic Concepts and Catalyst Design; Giménez, S., Bisquert, J., Eds.; Photoelectrochemical Solar Fuel Production; Springer: Cham, Switzerland, 2016. [Google Scholar]
  116. Suen, N.-T.; Hung, S.-F.; Quan, Q.; Zhang, N.; Xu, Y.-J.; Chen, H.M. Electrocatalysis for the oxygen evolution reaction: Recent development and future perspectives. Chem. Soc. Rev. 2017, 46, 337–365. [Google Scholar] [CrossRef] [PubMed]
  117. Reier, T.; Nong, H.N.; Teschner, D.; Schlögl, R.; Strasser, P. Electrocatalytic oxygen evolution reaction in acidic environments—Reaction mechanisms and catalysts. Adv. Energy Mater. 2017, 7, 1601275–1601293. [Google Scholar] [CrossRef]
  118. Matyushcov, V. Standard electrode potential, Tafel equation, and the solvation thermodynamics. J. Chem. Phys. 2009, 130, 234704–234804. [Google Scholar] [CrossRef] [PubMed]
  119. Trasatti, S. Electrocatalysis in the anodic evolution of oxygen and chlorine. Electrochim. Acta 1984, 29, 1503–1512. [Google Scholar] [CrossRef]
  120. Castelli, P.; Trasatti, S.; Pollak, F.H.; Ogrady, W.E. Single crystals as model electrocatalysts: Oxygen evolution on RuO2 (110). J. Electroanal. Chem. 1986, 210, 189–194. [Google Scholar] [CrossRef]
  121. Yoo, H.; Oh, K.; Lee, G.; Choi, J. RuO2-doped Anodic TiO2 nanotubes for water oxidation: Single-step anodization vs potential shock method. J. Electrochem. Soc. 2017, 164, H104–H111. [Google Scholar] [CrossRef]
  122. Bockris, J.O.; Nagy, Z. Symmetry factor and Transfer coefficient; A source of confusion in electrode kinetics. J. Chem. Edu. 1973, 50, 839–843. [Google Scholar] [CrossRef]
  123. Conway, B.E.; Salomon, M. Electrochemical reaction orders: Applications to the hydrogen- and oxygen-evolution reactions. Electrochim. Acta 1964, 9, 1599–1615. [Google Scholar] [CrossRef]
  124. Conway, B.E.; Bai, L.; Sattar, M.A. Role of the transfer coefficient in electrocatalysis: Applications to the H2 and O2 evolution reactions and the characterization of participating adsorbed intermediates. Int. J. Hydrogen Energy 1987, 12, 607–621. [Google Scholar] [CrossRef]
  125. Rossmeisl, J.; Logadottir, A.; Norskov, J.K. Electrolysis of water on (oxidized) metal surfaces. Chem. Phys. 2005, 319, 178–184. [Google Scholar] [CrossRef]
  126. Medford, A.J.; Voivodic, A.; Hummelshøj, J.S.; Voss, J.; Abild-Pedersen, F.; Studt, F.; Bligaard, T.; Nilsson, A.; Nørskov, J.K. From the Sabatier principle to a predictive theory of transition-metal heterogeneous catalysis. J. Catal. 2015, 328, 36–42. [Google Scholar] [CrossRef] [Green Version]
  127. Fabbri, E.; Havereder, A.; Waltar, K.; Kötz, R.; Schmidt, T.J. Developments and perspectives of oxide-based catalysts for the oxygen evolution reaction. Catal. Sci. Technol. 2014, 4, 3800–3821. [Google Scholar] [CrossRef] [Green Version]
  128. Kötz, R.; Stucki, S. Stabilization of RuO2 by IrO2 for anodic oxygen evolution in acid media. Electrochim. Acta 1986, 31, 1311–1316. [Google Scholar] [CrossRef]
  129. Yeo, R.S.; Orehotsky, J.; Visscher, W.; Srinivasan, S. Ruthenium-based mixed oxides as electrocatalysts for oxygen evolution in acid electrolytes. J. Electrochem. Soc. 1981, 128, 1900–1904. [Google Scholar] [CrossRef]
  130. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  131. Fujishima, A.; Zhang, X.; Tryk, D.A. TiO2 photocatalysis and related surface phenomena. Surf. Sci. Rep. 2008, 63, 515–582. [Google Scholar] [CrossRef]
  132. Masudy-Panah, S.; Eugene, Y.-J.K.; Khiavi, N.D.; Katal, R.; Gong, X. Aluminum-incorporated p-CuO/n-ZnO photocathode coated with nanocrystal-engineered TiO2 protective layer for photoelectrochemical water splitting and hydrogen generation. J. Mater. Chem. A 2018, 6, 11951–11965. [Google Scholar] [CrossRef]
  133. Katal, R.; Panah, S.M.; Zarinejad, M.; Salehi, M.; Jiangyong, H. Synthesis of self-gravity settling faceted-anatase TiO2 with dominant {010} facets for the photocatalytic degradation of acetaminophen and study of the type of generated oxygen vacancy in faceted-TiO2. Water 2018, 10, 1462. [Google Scholar] [CrossRef]
  134. Katal, R.; Salehi, M.; Farahani, M.H.D.A.; Masudy-Panah, S.; Ong, S.L.; Hu, J. Preparation of a new type of black TiO2 under a vacuum atmosphere for sunlight photocatalysis. ACS Appl. Mater. Interfaces 2018, 10, 35316–35326. [Google Scholar] [CrossRef] [PubMed]
  135. Moakhar, R.S.; Masudy-Panah, S.; Jalali, M.; Goh, G.K.L.; Dolati, A.; Ghorbani, M.; Riahi-Noori, N. Sunlight driven photoelectrochemical light-to-electricity conversion of screen-printed surface nanostructured TiO2 decorated with plasmonic Au nanoparticles. Electrochim. Acta 2016, 219, 386–393. [Google Scholar] [CrossRef]
  136. Ghosh, M.; Liu, J.; Chuang, S.S.C.; Jana, S.C. Fabrication of hierarchical V2O5 nanorods on TiO2 nanofibers and their enhanced photocatalytic activity under visible light. ChemCatChem 2018, 10, 3305–3318. [Google Scholar] [CrossRef]
  137. Paramasivam, I.; Macak, J.M.; Schmuki, P. Photocatalytic activity of TiO2 nanotube layers loaded with Ag and Au nanoparticles. Electrochem. Commun. 2008, 10, 71–75. [Google Scholar] [CrossRef]
  138. Macak, J.M.; Zlamal, M.; Krysa, J.; Schmuki, P. Self-organized TiO2 nanotube layers as highly efficient photocatalysts. Small 2007, 3, 300–304. [Google Scholar] [CrossRef] [PubMed]
  139. Sohn, Y.S.; Smith, Y.R.; Misra, M.; Subramanian, V. Electrochemically assisted photocatalytic degradation of methyl orange using anodized titanium dioxide nanotubes. Appl. Catal. B Environ. 2008, 84, 372–378. [Google Scholar] [CrossRef]
  140. Wang, D.; Liu, L.; Zhang, F.; Tao, K.; Pippel, E.; Domen, K. Spontaneous phase and morphology transformations of anodized titania nanotubes induced by water at room temperature. Nano Lett. 2011, 11, 3649–3655. [Google Scholar] [CrossRef] [PubMed]
  141. Paramasivam, I.; Nah, Y.-C.; Das, C.; Shrestha, N.K.; Schmuki, P. WO3/TiO2 nanotubes with strongly enhanced photocatalytic activity. Chem. Eur. J. 2010, 16, 8993–8997. [Google Scholar] [CrossRef] [PubMed]
  142. Mohapatra, S.K.; Misra, M.; Mahajan, V.K.; Raja, K.S. A novel method for the synthesis of titania nanotubes using sonoelectrochemical method and its application for photoelectrochemical splitting of water. J. Catal. 2007, 246, 362–369. [Google Scholar] [CrossRef]
  143. Allam, N.K.; Shankar, K.; Grimes, C.A. Photoelectrochemical and water photoelectrolysis properties of ordered TiO2 nanotubes fabricated by Ti anodization in fluoride-free HCl electrolytes. J. Mater. Chem. 2008, 18, 2341–2348. [Google Scholar] [CrossRef]
  144. Mohapatra, S.K.; Misra, M.; Mahajan, V.K.; Raja, K.S. Design of a highly efficient photoelectrolytic cell for hydrogen generation by water splitting:  Application of TiO2−xCx nanotubes as a photoanode and Pt/TiO2 nanotubes as a cathode. J. Phys. Chem. C 2007, 111, 8677–8685. [Google Scholar] [CrossRef]
  145. Mahajan, V.K.; Misra, M.; Raja, K.S.; Mohapatra, S.K. Self-organized TiO2 nanotubular arrays for photoelectrochemical hydrogen generation: Effect of crystallization and defect structure. J. Phys. D Appl. Phys. 2008, 41, 125307. [Google Scholar] [CrossRef]
  146. Nam, W.; Han, G.Y. Preparation and characterization of anodized Pt–TiO2 nanotube arrays for water splitting. J. Chem. Eng. Jpn. 2007, 40, 266–269. [Google Scholar] [CrossRef]
  147. Ghicov, A.; Schmidt, B.; Kunze, J.; Schmuki, P. Photoresponse in the visible range from Cr doped TiO2 nanotubes. Chem. Phys. Lett. 2007, 433, 323–326. [Google Scholar] [CrossRef]
  148. Das, C.; Roy, P.; Yang, M.; Jha, H.; Schmuki, P. Nb doped TiO2 nanotubes for enhanced photoelectrochemical water-splitting. Nanoscale 2011, 3, 3094–3096. [Google Scholar] [CrossRef] [PubMed]
  149. Roy, P.; Das, C.; Lee, K.; Hahn, R.; Ruff, T.; Moll, M.; Schmuki, P. Oxide nanotubes on Ti−Ru alloys: Strongly enhanced and stable photoelectrochemical activity for water splitting. J. Am. Chem. Soc. 2011, 133, 5629–5631. [Google Scholar] [CrossRef] [PubMed]
  150. Yang, M.; Jha, H.; Liu, N.; Schmuki, P. Increased photocurrent response in Nb-doped TiO2 nanotubes. J. Mater. Chem. 2011, 21, 15205–15208. [Google Scholar] [CrossRef]
  151. Altomare, M.; Lee, K.; Killian, M.S.; Selli, E.; Schmuki, P. Ta-Doped TiO2 nanotubes for enhanced solar-light photoelectrochemical water splitting. Chem. Eur. J. 2013, 19, 5841–5844. [Google Scholar] [CrossRef] [PubMed]
  152. Kawai, T.; Sakata, T. Conversion of carbohydrate into hydrogen fuel by a photocatalytic process. Nature 1980, 286, 474–476. [Google Scholar] [CrossRef]
  153. Sakata, T.; Hashimoto, K.; Kawai, T. Catalytic properties of ruthenium oxide on n-type semiconductors under illumination. J. Phys. Chem. 1984, 88, 5214–5221. [Google Scholar] [CrossRef]
  154. Blondeel, G.; Harriman, A.; Porter, G.; Urwin, D.; Kiwi, J. Design, preparation and characterization of ruthenium dioxide/titanium dioxide catalytic surfaces active in photooxidation of water. J. Phys. Chem. 1983, 87, 2629–2636. [Google Scholar] [CrossRef]
  155. O’Regan, B.; Grätzel, M. A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 1991, 353, 737–740. [Google Scholar] [CrossRef]
  156. Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. [Google Scholar] [CrossRef] [PubMed]
  157. Nazeeruddin, M.K.; Angelis, F.D.; Fantacci, S.; Selloni, A.; Viscardi, G.; Liska, P.; Ito, S.; Takeru, B.; Grätzel, M. Combined experimental and DFT-TDDFT computational study of photoelectrochemical cell ruthenium sensitizers. J. Am. Chem. Soc. 2005, 127, 16835–16847. [Google Scholar] [CrossRef] [PubMed]
  158. Tachibana, Y.; Moser, J.E.; Grätzel, M.; Klug, D.R.; Durrant, J.R. Subpicosecond interfacial charge separation in dye-sensitized nanocrystalline titanium dioxide films. J. Phys. Chem. 1996, 100, 20056–20062. [Google Scholar] [CrossRef]
  159. Peter, L.M. Characterization and modeling of dye-sensitized solar cells. J. Phys. Chem. C 2007, 111, 6601–6612. [Google Scholar] [CrossRef]
  160. Bailes, M.; Cameron, P.J.; Lobato, K.; Peter, L.M. Determination of the density and energetic distribution of electron traps in dye-sensitized nanocrystalline solar cells. J. Phys. Chem. B 2005, 109, 15429–15435. [Google Scholar] [CrossRef] [PubMed]
  161. Cameron, P.J.; Peter, L.M. How does back-reaction at the conducting glass substrate influence the dynamic photovoltage response of nanocrystalline dye-sensitized solar cells? J. Phys. Chem. B 2005, 109, 7392–7398. [Google Scholar] [CrossRef] [PubMed]
  162. Jennings, J.R.; Peter, L.M. A Reappraisal of the electron diffusion length in solid-state dye-sensitized solar cells. J. Phys. Chem. C 2007, 111, 16100–16104. [Google Scholar] [CrossRef]
  163. Fujihara, K.; Kumar, A.; Jose, R.; Ramakrishna, S.; Uchida, S. Spray deposition of electrospun TiO2 nanorods for dye-sensitized solar cell. Nanotechnology 2007, 18, 365709. [Google Scholar] [CrossRef]
  164. Zhu, K.; Vinzant, T.B.; Neale, N.R.; Frank, A.J. Removing structural disorder from oriented tio2 nanotube arrays:  Reducing the dimensionality of transport and recombination in dye-sensitized solar cells. Nano Lett. 2007, 7, 3739–3746. [Google Scholar] [CrossRef] [PubMed]
  165. Zhu, K.; Neale, N.R.; Miedaner, A.; Frank, A.J. Enhanced charge-collection efficiencies and light scattering in dye-sensitized solar cells using oriented TiO2 nanotubes arrays. Nano Lett. 2007, 7, 69–74. [Google Scholar] [CrossRef] [PubMed]
  166. Jennings, J.R.; Ghicov, A.; Peter, L.M.; Schmuki, P.; Walker, A.B. Dye-sensitized solar cells based on oriented TiO2 nanotube arrays: Transport, trapping, and transfer of electrons. J. Am. Chem. Soc. 2008, 130, 13364–13372. [Google Scholar] [CrossRef] [PubMed]
  167. Zhang, W.; Xie, Y.; Xiong, D.; Zeng, X.; Li, Z.; Wang, M.; Cheng, Y.-B.; Chen, W.; Yang, K.; Yang, S. TiO2 nanorods: A facile size- and shape-tunable synthesis and effective improvement of charge collection kinetics for dye-sensitized solar cells. ACS Appl. Mater. Interfaces 2014, 6, 9698–9704. [Google Scholar] [CrossRef] [PubMed]
  168. Yang, M.; Kim, D.; Jha, H.; Lee, K.; Paul, J.; Schmuki, P. Nb doping of TiO2 nanotubes for an enhanced efficiency of dye-sensitized solar cells. Chem. Commun. 2011, 47, 2032–2034. [Google Scholar] [CrossRef] [PubMed]
  169. Lee, K.; Schmuki, P. Ta doping for an enhanced efficiency of TiO2 nanotube based dye-sensitized solar cells. Electrochem. Commun. 2012, 25, 11–14. [Google Scholar] [CrossRef]
Figure 1. Schematic of the reaction mechanism of TiO2 anodization. (a) O2 and Ti4+ are transported through the oxide layers due to the electric field. (b) In presence of fluoride ions, [TiF6]2 and an F-rich layer are generated. The initial formation of a tubular structure is shown in (c), and in (d) and (e) the F-rich layer is finally dissolved by the water in the electrolyte. Reproduced with permission from ref [11]. Friedrich-Alexander-Universität Erlangen-Nürnberg (2013).
Figure 1. Schematic of the reaction mechanism of TiO2 anodization. (a) O2 and Ti4+ are transported through the oxide layers due to the electric field. (b) In presence of fluoride ions, [TiF6]2 and an F-rich layer are generated. The initial formation of a tubular structure is shown in (c), and in (d) and (e) the F-rich layer is finally dissolved by the water in the electrolyte. Reproduced with permission from ref [11]. Friedrich-Alexander-Universität Erlangen-Nürnberg (2013).
Catalysts 08 00555 g001
Figure 2. A representative current–time curve for the formation of tubular structures. (I) The current decreases drastically due to the growth of the barrier oxide during initial stage of anodization. Next, (II) nano-pits or cracks caused by F ions cause an increase in the current due to the concentrated electric fields on these pits. Once the tubular structure has formed and is growing evenly, the current declines and converges at a stable value due to the increased diffusion length of the ions involved in the reaction. Reproduced with permission from ref [11]. Friedrich-Alexander-Universität Erlangen-Nürnberg (2013).
Figure 2. A representative current–time curve for the formation of tubular structures. (I) The current decreases drastically due to the growth of the barrier oxide during initial stage of anodization. Next, (II) nano-pits or cracks caused by F ions cause an increase in the current due to the concentrated electric fields on these pits. Once the tubular structure has formed and is growing evenly, the current declines and converges at a stable value due to the increased diffusion length of the ions involved in the reaction. Reproduced with permission from ref [11]. Friedrich-Alexander-Universität Erlangen-Nürnberg (2013).
Catalysts 08 00555 g002
Figure 3. (a) A schematic of the conventional anodization procedure and doping methods. Note that potential shock is a sequential two-step anodic process. (b) Left: The doping of the catalysts follows “plastic flow.” The doped catalysts are able to partially dissolve at long doping times, as the high anodic potential induces the oxygen evolution reaction. Right: The morphologies of the doped TiO2 nanotubes depends on the doping method. Reproduced with permission from ref [31]. Inha University (2018).
Figure 3. (a) A schematic of the conventional anodization procedure and doping methods. Note that potential shock is a sequential two-step anodic process. (b) Left: The doping of the catalysts follows “plastic flow.” The doped catalysts are able to partially dissolve at long doping times, as the high anodic potential induces the oxygen evolution reaction. Right: The morphologies of the doped TiO2 nanotubes depends on the doping method. Reproduced with permission from ref [31]. Inha University (2018).
Catalysts 08 00555 g003
Figure 4. (a) TiO2 nanotubes grown conventionally in ethylene glycol. (b) The thickness of the TiO2 nanotubes is almost doubled by adding RuO4 (as KRuO4) to the electrolyte during single-step anodization. (c) The thickness of the TiO2 nanotubes doped with RuO2 is further increased by the addition of water to the RuO2 precursor-containing electrolyte. (d) After doping via single-step anodization, a RuO2 rutile phase is clearly observed in the TiO2 anatase phase. Reproduced with permission from ref [17]. WILEY-VCH Verlag GmbH and Co. KGaA, Weinheim (2015).
Figure 4. (a) TiO2 nanotubes grown conventionally in ethylene glycol. (b) The thickness of the TiO2 nanotubes is almost doubled by adding RuO4 (as KRuO4) to the electrolyte during single-step anodization. (c) The thickness of the TiO2 nanotubes doped with RuO2 is further increased by the addition of water to the RuO2 precursor-containing electrolyte. (d) After doping via single-step anodization, a RuO2 rutile phase is clearly observed in the TiO2 anatase phase. Reproduced with permission from ref [17]. WILEY-VCH Verlag GmbH and Co. KGaA, Weinheim (2015).
Catalysts 08 00555 g004
Figure 5. Cross-sectional scanning electron microscopy (SEM) images of RuO2-doped TiO2 nanotubes prepared by the potential shock method. At the bottom of the tubes, additional barrier oxide was formed; the thickness of the barrier oxide depended on the applied shock voltage.
Figure 5. Cross-sectional scanning electron microscopy (SEM) images of RuO2-doped TiO2 nanotubes prepared by the potential shock method. At the bottom of the tubes, additional barrier oxide was formed; the thickness of the barrier oxide depended on the applied shock voltage.
Catalysts 08 00555 g005
Figure 6. Representative cross-sectional SEM images of (a,b) regular and (cf) irregular nanotubes formed by the anodization of Ti alloys. The nanotubes were formed by the anodization of Ti alloys containing (a,b) 0.02 at.%, (c,d) 0.05 at.%, and (e,f) 0.2 at.% of Ru. Reproduced with permission from reference [114]. WILEY-VCH Verlag GmbH and Co. KGaA, Weinheim (2012).
Figure 6. Representative cross-sectional SEM images of (a,b) regular and (cf) irregular nanotubes formed by the anodization of Ti alloys. The nanotubes were formed by the anodization of Ti alloys containing (a,b) 0.02 at.%, (c,d) 0.05 at.%, and (e,f) 0.2 at.% of Ru. Reproduced with permission from reference [114]. WILEY-VCH Verlag GmbH and Co. KGaA, Weinheim (2012).
Catalysts 08 00555 g006
Figure 7. Linear relationship of the alkaline OER current density with the overpotential and the morphological parameters of doping concentration, surface area, and nanotube thickness for RuO2-doped TiO2 nanotubes. ①, ②, and ③ correspond to RuO2 doping via single-step anodization in organic solvent, single-step anodization in aqueous conditions, and potential shock in aqueous conditions. Reproduced with permission from reference [31]. Inha University (2018).
Figure 7. Linear relationship of the alkaline OER current density with the overpotential and the morphological parameters of doping concentration, surface area, and nanotube thickness for RuO2-doped TiO2 nanotubes. ①, ②, and ③ correspond to RuO2 doping via single-step anodization in organic solvent, single-step anodization in aqueous conditions, and potential shock in aqueous conditions. Reproduced with permission from reference [31]. Inha University (2018).
Catalysts 08 00555 g007
Figure 8. Morphological characterization data of IrO2-RuO2-doped TiO2 nanotubes and the corresponding OER performance results. The elements (a) Ru (red spot) and Ir (green spot) are evenly distributed throughout the TiO2. (b) The nanotubes are polycrystalline, with the rutile phase of RuO2 and IrO2 well-dispersed in TiO2 anatase phase. (c) LSV graphs showing overpotentials and gas chromatography mass spectroscopy (GC-MS) results measured for different nanotube materials in identical alkaline OER tests. Reproduced with permission from reference [31]. Inha University (2018).
Figure 8. Morphological characterization data of IrO2-RuO2-doped TiO2 nanotubes and the corresponding OER performance results. The elements (a) Ru (red spot) and Ir (green spot) are evenly distributed throughout the TiO2. (b) The nanotubes are polycrystalline, with the rutile phase of RuO2 and IrO2 well-dispersed in TiO2 anatase phase. (c) LSV graphs showing overpotentials and gas chromatography mass spectroscopy (GC-MS) results measured for different nanotube materials in identical alkaline OER tests. Reproduced with permission from reference [31]. Inha University (2018).
Catalysts 08 00555 g008
Figure 9. Schematic diagram of a dye-sensitized solar cell (DSSC): (a) Principle of a dye-sensitized solar cell and the time scales of various processes; (b) a full cell constructed with anodic TiO2 nanotubes. Reproduced with permission from reference [8]. American Chemical Society (2014).
Figure 9. Schematic diagram of a dye-sensitized solar cell (DSSC): (a) Principle of a dye-sensitized solar cell and the time scales of various processes; (b) a full cell constructed with anodic TiO2 nanotubes. Reproduced with permission from reference [8]. American Chemical Society (2014).
Catalysts 08 00555 g009
Figure 10. (a) Recombination time (τr) and (b) electron transfer (τc) constants calculated from IMVS and IMPS measurements for pure TiO2 and 0.1 Nb–TiO2 nanotube-layer-based DSSCs. (c) Electrochemical impedance measurements of TiO2 and Ru-doped TiO2 nanotubes in 0.1 M Na2SO4. Reproduced with permission from ref [168]. Royal Society of Chemistry (2010) and reference [114]. WILEY-VCH Verlag GmbH and Co. KGaA, Weinheim (2012).
Figure 10. (a) Recombination time (τr) and (b) electron transfer (τc) constants calculated from IMVS and IMPS measurements for pure TiO2 and 0.1 Nb–TiO2 nanotube-layer-based DSSCs. (c) Electrochemical impedance measurements of TiO2 and Ru-doped TiO2 nanotubes in 0.1 M Na2SO4. Reproduced with permission from ref [168]. Royal Society of Chemistry (2010) and reference [114]. WILEY-VCH Verlag GmbH and Co. KGaA, Weinheim (2012).
Catalysts 08 00555 g010

Share and Cite

MDPI and ACS Style

Yoo, H.; Kim, M.; Kim, Y.-T.; Lee, K.; Choi, J. Catalyst-Doped Anodic TiO2 Nanotubes: Binder-Free Electrodes for (Photo)Electrochemical Reactions. Catalysts 2018, 8, 555. https://doi.org/10.3390/catal8110555

AMA Style

Yoo H, Kim M, Kim Y-T, Lee K, Choi J. Catalyst-Doped Anodic TiO2 Nanotubes: Binder-Free Electrodes for (Photo)Electrochemical Reactions. Catalysts. 2018; 8(11):555. https://doi.org/10.3390/catal8110555

Chicago/Turabian Style

Yoo, Hyeonseok, Moonsu Kim, Yong-Tae Kim, Kiyoung Lee, and Jinsub Choi. 2018. "Catalyst-Doped Anodic TiO2 Nanotubes: Binder-Free Electrodes for (Photo)Electrochemical Reactions" Catalysts 8, no. 11: 555. https://doi.org/10.3390/catal8110555

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