You are on page 1of 16

Applied Catalysis A: General 481 (2014) 127–142

Contents lists available at ScienceDirect

Applied Catalysis A: General


journal homepage: www.elsevier.com/locate/apcata

Review

A critical review on the recent progress of synthesizing techniques


and fabrication of TiO2 -based nanotubes photocatalysts
Yean Ling Pang ∗ , Steven Lim, Hwai Chyuan Ong, Wen Tong Chong
Department of Mechanical Engineering, Faculty of Engineering, University of Malaya, 50603 Kuala Lumpur, Malaysia

a r t i c l e i n f o a b s t r a c t

Article history: One-dimensional titanium dioxide (TiO2 )-based nanotubes have attracted great interest to be used as
Received 28 February 2014 photocatalysts in the field of environmental applications. Three main approaches via template-assisted,
Received in revised form 13 April 2014 electrochemical anodic oxidation and alkaline hydrothermal treatment used to prepare the nanostruc-
Accepted 12 May 2014
tured TiO2 -based nanotubes are reviewed. The parameters that affect the formation of TiO2 -based
Available online 20 May 2014
nanotubes via hydrothermal method such as phases and particle sizes of starting materials, types and con-
centrations of alkaline solution, temperature and duration of hydrothermal treatment, ultrasonication-
Keywords:
and microwave-assisted hydrothermal synthesis, acid washing and calcination have been reviewed in
TiO2 nanotubes
Photocatalyst
details. This paper also discussed the possible crystal structure and formation mechanism of TiO2 -based
Hydrothermal synthesis nanotubes via alkaline hydrothermal treatment. In addition, the recent research progress on the struc-
Mechanism tural modification of TiO2 -based nanotubes to be used as photocatalysts is summarized in this review.
Modifications Modification strategies investigated include metal nanoparticles deposition, single- and co-doping of
metal ions/non-metal ions, coupled with other semiconductors to form binary composites and hybrid
with carbon nanomaterials.
© 2014 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
2. TiO2 nanoparticles, randomized and oriented nanotubes for photocatalytic reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
3. Synthesizing techniques of TiO2 nanotubes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
3.1. Template-assisted method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
3.2. Electrochemical anodization method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
3.3. Hydrothermal method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
4. Factors influencing the formation of titanate nanotubes via hydrothermal treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
4.1. Starting materials: phases and particle sizes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
4.2. Alkaline solution: types and concentrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
4.3. Hydrothermal treatment: temperature and duration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
4.4. Assisted hydrothermal synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
4.5. Post-hydrothermal treatments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
4.5.1. Acid washing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
4.5.2. Calcination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
5. Crystal structure, formation mechanism and structural-activity of titanate nanotubes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
6. Modification of titanate nanotubes as photocatalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
6.1. Metal-loaded . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
6.2. Doping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
6.3. Coupled semiconductors/binary composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
6.4. Hybrid with carbon nanomaterials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

∗ Corresponding author. Tel.: +60 16 408 0412; fax: +60 3 7967 5317.
E-mail address: pangyeanling@hotmail.com (Y.L. Pang).

http://dx.doi.org/10.1016/j.apcata.2014.05.007
0926-860X/© 2014 Elsevier B.V. All rights reserved.
128 Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142

7. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

1. Introduction nanotubes photocatalyts through hydrothermal treatment has


been emphasized in this review. The factors during the hydro-
Titanium dioxide (TiO2 ) is a very well-known photocatalysis thermal reaction which affect the formation of TiO2 -based
material due to its superior photo-reactivity, non-toxicity, long nanotubes, possible mechanism, modification of TiO2 -based nano-
term stability, high corrosion resistance and available in low cost tubes and their effects on the photocatalytic reactions have been
[1–3]. The key to photolysis of water (H2 O) molecule under illumi- reviewed in details.
nation of light is the positions of the conduction and valence bands
relative to the redox potential of water [4]. TiO2 fulfills the par- 2. TiO2 nanoparticles, randomized and oriented nanotubes
ticular requirement as compared to the other metal oxides, where for photocatalytic reactions
the conduction band edge lay higher than that of the species to
be reduced (hydrogen ion/hydrogen) and the valence band edge For better clarification, the mechanism of photocatalytic reac-
located lower than that of the species to be oxidized (H2 O/oxygen tion is briefly summarized in Eqs. (1)–(6) [13,15]. Photocatalytic
(O2 )). process occurs based on the sufficient energy obtained from
The drawbacks of TiO2 nanoparticles photocatalyst include absorption of light radiation which equals to or is greater than the
the requirement of large amount of catalysts, difficulty in cata- semiconductor band gap energy. This will lead to the excitation and
lyst recycling and problematic agglomeration of TiO2 nanocrystals transfer of electrons from the valence band to the conduction band,
into large particles [5]. Although immobilized TiO2 is a feasible leaving behind holes at the valence band. The electrons could react
option, the overall photocatalytic activity has been compromised with electron acceptors such as O2 adsorbed on the catalyst surface
due to the reduction of surface area and limitation in mass trans- or dissolved in water and reduce to superoxide radical anion (• O2 − ).
fer [6]. In order to enhance photocatalytic activity and facilitate Meanwhile, holes tend to oxidize directly on the organic molecules
catalyst separation after photocatalytic reaction, one-dimensional or react with hydroxyl group and oxidize to hydroxyl radical (• OH).
(1D) nanostructured materials such as nanotubes, nanowires and The resulting radicals are capable of promoting oxidation reactions
nanorods have been investigated. Besides, charge recombination to mineralize organic pollutant.
probability within TiO2 nanoparticles also increases due to the
presence of defects, surface states and grain boundaries which act TiO2 + energy(heatorlight) → hvb + + ecb − (1)
as electron trapping sites [7]. Meanwhile, 1D nanostructured mate- − •O −
O2(ads) + ecb → 2 (2)
rials are expected to have faster electron transport and lower charge
+ • OH +
recombination due to 1D channel for electron transportation and H2 O + hvb → +H (3)
decrement of inter-crystalline contacts, respectively. +
hvb + pollutant → • pollutant+ → degradationproducts (4)
The present review focuses on TiO2 -based nanotubes rather
•O − + pollutant → degradationproducts + CO2 (5)
than nanowires and nanorods due to the internal layer of the hollow 2
nanotubes provide larger surface area and easily functionalized by • OH + pollutant → degradationproducts + CO2 (6)
filling the hollow nanotubes [8]. However, TiO2 is a wide band gap
semiconductor with 3.2 eV for anatase [1,9], 3.0 eV for rutile [1,9] + −
where hvb and ecb are the positive hole in the valence band and
and nanotubes material is reported to have a mobility band gap the electron in the conduction band, respectively while subscripts
of about 3.1–3.87 eV [10,11]. The blue shift of the absorption edge ‘ads’ denote the adsorbed forms of the O2 molecule.
wavelength is attributed to the quantum size effect of TiO2 nano- In randomly packed TiO2 nanoparticles, the transportation of
tubes, which is due to the thin nanotube wall thickness of around photo-excited electron is limited due to low electron diffusion
1 nm [12]. This is the major drawback that limits the use of TiO2 coefficient and scattering effect of free electrons [7]. The electron
photocatalyts for large scale application. Therefore, intense effort trapped sites could occur at defects, surface states and grain bound-
focuses on the modification of TiO2 nanotubes by various strate- aries between nanoparticles [16]. Consequently, charge collection
gies such as metal nanoparticles deposition, single or co-doping efficiency was diminished due to the increment of charge recom-
metal ions/non-metal ions, coupling with other semiconductors bination at the trapping sites. Moreover, other drawbacks of TiO2
and hybrid with carbon nanomaterials. These efforts alter the elec- nanoparticles are often related to the separation and recycling of
tronic properties of TiO2 nanotubes by narrowing the band gap the particulate catalysts from the reaction media [5]. Filtration of
energy, resulting in higher reaction rates for the photocatalytic ultra-fine nanoparticles is a tedious and time consuming separation
degradation of organic pollutants. process.
The review begins with a brief overview of the synthesis In order to improve these behaviors, recent researches have
techniques including template-assisted, electrochemical anodiza- moved toward the use of 1D nanostructures of TiO2 such as nano-
tion and hydrothermal treatment with their research progress, tubes, nanorods and nanowires. The unique properties of high
advantages and disadvantages. Extensive applications of template- aspect ratio of TiO2 nanotubes include large surface area, high
assisted method often encounter difficulties like pore size cation exchangeability, high catalytic activity, easier separation
limitation of the prepared template and post-removal template and recyclability [8]. These advantages make nanotubes attrac-
usually results in impurities [13]. The method of electrochemi- tive for industrial scale applications. Ohsaki et al. [17] found that
cal anodization is based on the anodization of titanium foil to the use of TiO2 nanotubes, even in a randomly disordered system,
obtain nanoporous TiO2 film [14] and can be considered as an was able to suppress the possibility of charge recombination. At
expensive technique for large scale production of TiO2 nanotubes the same electron diffusion coefficient of TiO2 nanoparticles and
[3]. Besides, this method may lead to un-separatable tubes and randomly packed TiO2 nanotubes, TiO2 nanotubes with longer dif-
uneven length distribution of tubes over a large surface area [9,10]. fusion length and longer electron lifetime would exhibit excellent
Thus, the recent research progress of the formation of TiO2 -based electron transport. In addition, TiO2 nanotubes can provide strong
Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142 129

Fig. 1. A comparison of the electron pathways through (a) nanoparticles [16], (b) randomized [18] and (c) oriented nanotubular structured TiO2 [16].

light scattering effect and enhance light harvesting properties [16]. Table 1
Comparison of available methods for TiO2 nanotubes preparation [19].
The special features of internal and external surface area are avail-
able for the adsorption and chemical reaction of organic pollutant Fabrication method Characteristics
molecules. Template-assisted • Ordered arrays
It should be highlighted that TiO2 nanotubes can be synthesized method • Advantages:
into randomized powdery or immobilized nanotubes arrays. The The dimension of nanotubes can be controlled by the
immobilized nanotubes are self-organized, vertically oriented and size and type of applied templates
Uniform size of nanotubes can be formed
supported on a surface to form an integrated unit. Immobilization
• Disadvantages:
of nanocatalyst on a support can prevent catalyst agglomeration Nanotubes morphology may be destroyed during
and facilitate catalyst recycling. The arrangement of highly ordered post-removal of the templates
TiO2 nanotubes array is perpendicular to the titanium substrate, Dissolution of template may result in contamination of
nanotubes
providing channels for efficient charge transfer from the solution
to the collecting substrate [7]. This reduces the losses incurred by Electrochemical • Oriented arrays
charge trapping across the nanoparticles grain boundaries [16]. anodization method • Advantages:
The dimension of nanotubes can be controlled by
Fig. 1 shows the electron transporting pathways in nanoparti-
varying the voltage, electrolyte, pH and anodizing time
cles, randomized and oriented nanotubular structured TiO2 . The Ordered alignment of nanotubes with high aspect ratio
ordered TiO2 nanotubes arrays are expected to improve the charge can be formed
collection efficiency due to the fast electron transportation along 1D • Disadvantages:
The requirement of fabrication apparatus
channel with minimum charge recombination site. Consequently,
Length distribution and separation of nanotubes over a
a lot of research works had been concentrating on the formation of large surface area is not well-developed
ordered arrays of nanotubes especially in the application of pho-
Hydrothermal • Random alignment or can be aligned
toelectrochemical water splitting and dye sensitized solar cells
treatment • Advantages:
[7,9]. However, dispersed non-ordered nanotubes showed great Easy route to obtain nanotubes in relatively large
potential as heterogeneous catalyst to accelerate photocatalytic amount
degradation of organic pollutants in liquid medium [18]. Thus, the • Disadvantages:
research progresses on the formation of nanotubes via different Difficult in achieving uniform size of nanotubes

synthesizing techniques are discussed in details in subsequent sec-


tion.
3.1. Template-assisted method

3. Synthesizing techniques of TiO2 nanotubes The dimensions (diameter and length) of TiO2 nanotubes formed
through hard template approach membranes can be controlled eas-
Three popular synthesizing techniques of TiO2 nanotubes have ily by the dimension of templates [20]. Generally, TiO2 nanotubes
been investigated in recent years including template-assisted, derived from the templates are obtained after removal of tem-
electrochemical anodization and hydrothermal treatment. A com- plates by selective chemical etching [21] or thermal decomposition
parison between unique features of these three fabrication [22]. Template can be divided into two types: positive and nega-
methods is summarized in Table 1. tive templates as shown in Fig. 2 [20]. Positive template synthesis

Fig. 2. Illustration of two different types of hard templates: (a) positive template and (b) negative template [23].
130 Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142

Table 2
Template-assisted method for preparation of TiO2 nanotubes.

Template categories Types of Preparation methods of Dimensions of nanotubes References


(i) precursors (i) template
(ii) template (fabrication time) (ii) TiO2 nanotubes

Positive template (i) Amorphous TiO2 (i) Two-step anodization Internal diameter (Di ) = 70–100 nm Hoyer [24]
(ii) Anodized aluminum oxide (AAO) (ii) Template-based electrochemical Outer diameter (Do ) = 140–180 nm
(>3 h) acted as template for PMMA rods deposition process Average length (L) = 8 ␮m
(i) TiO2 (i) Low-temperature aqueous Wall thickness = 50–120 nm for Lee et al. [25]
(ii) Zinc oxide (ZnO) nanorod template solution method deposition times of 3–10 h
(5 h) (ii) Template-based liquid phase
deposition
(i) TiO2 sol (i) Sol–gel followed by Average diameter (D) = 400–600 nm Qiu et al. [21]
(ii) ZnO nanorod template (no fabrication hydrothermal method L = 30 ␮m
time is reported) (ii) Layer-by-layer absorption and Shell thickness = 150 nm
reaction technique Do and L of TiO2 branches were 60 nm
and 1 ␮m, respectively
(i) TiO2 sol (i) Rapid thermal chemical vapor Do = 60 nm Kim et al. [22]
(ii) Multi-walled carbon nanotubes (>2 h) deposition Wall thickness = 15 nm
(ii) Sol–gel method

Negative template (i) Titanium (IV) iso-propoxide (i) Commercially available Do = 80–100 nm Bae et al. [26]
(ii) AAO (ii) Template-based gas phase L = 30 ␮m
atomic layer deposition Wall thickness = 2.5 nm
Wall thickness = 7 nm
Specific surface area = 400 m2 /g
(i) Titanium (IV) t-butoxide (i) Commercially available Do = 200 nm Yuan et al. [27]
(ii) AAO (ii) Template-based alcoholate L = 60 ␮m
hydrolysis Wall thickness = 10–45 nm at
precursors concentration of 0.1–2.0 M

is employed when oxide materials are coated on the outer sur- three-dimensional (3D) branched TiO2 nanotube arrays on trans-
face of the template [21]. Meanwhile, negative template synthesis parent conducting oxide substrates using ZnO nanorods and ZnO
involved oxide materials which are deposited inside the template spacers as template. The ZnO nanorods template and sacrificial
channel space [23]. Table 2 shows several selected research works spacers can be removed by selective chemical etching in dilute
on the production of TiO2 nanotubes through positive and negative aqueous solution of titanium tetrachloride, which might lead to
templates. the collapse of TiO2 tubes.
In 1996, Hoyer [24] was the first researcher reported the Recently, carbon nanotubes have been considered to be an ideal
preparation of TiO2 nanotubes via template-assisted method. TiO2 template to synthesis TiO2 nanotubes due to the template easily
nanotubes were formed by electrochemical deposition method being removed, small diameter of carbon nanotubes and can sup-
into polymethyl methacrylate (PMMA) mold. Anodically grown port tubular morphology [22]. A sintering process at an appropriate
of porous anodized aluminum oxide (AAO) membrane consists temperature is performed after coating TiO2 on carbon nanotubes
of an array of parallel straight nanopores with uniform diam- to burned out the carbon nanotubes template and high crystallinity
eter and length. It served as negative template to produce a of TiO2 nanotubes could be obtained [22]. Meng et al. [32] reported
positive type polymer mold. TiO2 nanotubes were obtained after crystalline anatase TiO2 nanotubes were prepared using atomic
dissolution of the polymer mold using acetone. Recently, Kara- layer deposition on carbon nanotubes and calcination at 250 ◦ C.
man et al. [28] demonstrated that TiO2 nanotubes thin films Meanwhile, Kim et al. [22] successfully synthesized anatase TiO2
could be deposited over electrospun PMMA fibers by hot fila- nanotubes by performing a simply calcination process above 500 ◦ C
ment chemical vapor deposition method. Post-heat treatment of on TiO2 coated multi-walled carbon nanotubes composite materi-
the deposited materials leads to the decomposition of polymeric als via sol–gel method.
inner layer and formation of randomly distributed anatase TiO2 The advantages of positive template are inner diameters and
nanotubes. Several types of soft template had been reported in the length of the resulting nanotubes could be determined by the
literature such as core–shell–corona cylindrical polymer brushes, outer diameters and length of the original template, whereas outer
self-assembled block copolymers and biological superstructures diameters are determined by the thickness of deposited wall lay-
[29]. Ku et al. [30] introduced silicon moiety into a block copolymer, ers [20]. Meanwhile, the disadvantage of this type of template is
polystyrene-block-poly(4-(tert-butyldimethylsilyl)oxystyrene) to high difficulty in fabrication of uniform length and open ended
enhance template properties such as thermal stability, dielectric- nanotubes. Thereby, the use of negative template has attracted
ity and mechanical strength. They found that the wall thickness of attention.
the TiO2 nanotubes can be easily tuned by adjusting the number of Nanoporous AAO, which prepared by electrochemical oxidation
atomic layer. of aluminum in acidic solution has been extensively used as a nega-
On the other hand, zinc oxide (ZnO) nanostructure was tive template. The attractive advantages of anodic AAO membranes
employed as a template due to the low material cost and ease are large specific surface area, ordered pore geometry, adjustable
of fabrication [31]. It has characteristic of amphoteric in nature, structured properties such as pore diameter and length and the
which can be easily dissolved in mild acids or bases [31]. Lee et al. ease of dissolution with chemical solvents [23]. Assisted by AAO
[25] demonstrated that TiO2 nanotubes could be produced with- templates, various methods have been developed to prepare TiO2
out using chemical medium to dissolve ZnO nanorod template. The nanotubes such as electrochemical technique, sol–gel and atomic
removal of ZnO could be achieved by the reaction with hydrogen layer deposition [27]. By controlling the conditions of anodization
ions during liquid phase deposition process. However, the sam- process, 10–400 nm of pore diameter, 50–600 nm of inter-pore dis-
ple could be contaminated easily by precipitation resulting from tance, pore aspect ratio of 10 to 5000 and 10 nm to 150 ␮m of
the liquid phase deposition process of TiO2 . Qiu et al. [21] formed thickness of porous aluminum oxide layer can be prepared [23].
Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142 131

Table 3
Electrochemical anodization method for preparation of TiO2 nanotubes.

Anodization categories Types of Conditions anodization Dimensions of References


(i) titanium nanotubes
(ii) electrolytes

Hydrofluoric acid (HF) based (i) Titanium foil (thickness: n.d.) Voltages: 10–40 V D = 60 nm Gong et al. [33]
electrolytes (first generation) (ii) 0.5–3.5 wt.% HF aqueous solution Duration: 20 min to 6 h L = 250 nm
(i) Titanium foil (thickness: 0.2 mm) Voltages: 20 V D = 40–110 nm Şennik et al. [35]
(ii) 1 wt.% HF aqueous solution Duration: 45 min L = 1 ␮m

Water-based electrolyte containing (i) Titanium foil (thickness: 0.2 mm) Voltages: 20 V D = 90 nm Sreekantan et al. [34]
fluoride ion (second generation) (ii) 3 wt.% of ammonium fluoride (NH4 F) in 1 M Duration: 30 min to 3 h L = 0.7–2.5 ␮m
sodium sulfate (Na2 SO4 ) solution Wall thick-
ness = 20 nm
(i) Titanium foil (thickness: 0.5 mm) Voltages: 20 V Pore diame- Baram and Ein-Eli [36]
(ii) 0.5 wt.% sodium fluoride in 1 M Na2 SO4 Duration: 2 h ter = 75 nm
Surface
area = 94 cm2
on 4 cm2
sample

Organic solvent electrolyte containing (i) Titanium plate (thickness: n.d.) Voltages: 5–40 V Di = 15–53 nm Roman et al. [37]
fluoride ion (third generation) (ii) 0.55 wt.% NH4 F, 1% H2 O in ethylene glycerol Duration: 15–300 min Do = 45–67 nm
or 0.7 wt.% NH4 F, 9.3% H2 O in glycerol L = 1–3 ␮m
Wall thick-
ness = 7 nm
(i) Titanium foil (thickness: 0.13 mm) Voltages: 10–40 V D = 30–219 nm Nischk et al. [38]
(ii) 0.09 M NH4 F and 2% (v/v) H2 O in glycol or Duration: 1 h L = 0.5–5.9 ␮m
0.27 M NH4 F and 45% (v/v) H2 O in ethylene Wall thick-
glycol ness = 10–25 nm

Other approach (i) Titanium foil (thickness: 0.5 mm) Voltages: 10 V D = 7.53 ± 1.53 nm Antony et al. [14]
(ii) 0.1 M perchloric acid Duration: n.d. L = 25 ␮m
(i) Titanium foil (thickness: 100 ␮m) Voltages: 40 V D = 20–30 nm Jha et al. [39]
(ii) 0.05 M sodium perchlorate and 0.05 M Duration: 30 min L = >220 ␮m
sodium chloride in water–ethanol mixed followed 20 min
solution

Note: n.d. is defined as no data provided.

This makes the anodic AAO membrane as an excellent template can be divided into three generations based on the length of TiO2
material for 1D of ordered arrays nanostructures. nanotubes or type of electrolyte as shown in Table 3.
Bae et al. [26] successfully fabricated multi-walled anatase TiO2 In 1999, Zwilling et al. [40] had reported that the titanium
nanotubes using AAO as a negative template. Multi-walled nano- metal surface became porous when using anodization in chromic
tubes were produced by alternating TiO2 and aluminum oxide acid electrolytes containing hydrofluoric acid. They recognized that
nanolaminate structures onto porous aluminum oxide templates small addition of fluoride ions (F− ) to electrolyte was a crucial
with atomic layer deposition method, followed by wet etching of ingredient in the formation of self-organized porous structures.
the sacrificial aluminum oxide. The structural parameters including The latter work done by Gong et al. [33] demonstrated clearly self-
diameter, length, wall thickness, inter-wall spacing and number of organized and highly uniform nanotubes on titanium substrate
wall layers of the multi-walled TiO2 nanotubes can be adjusted. In using an aqueous dilute hydrofluoric solution as electrolyte. The
addition, Yuan et al. [27] synthesized TiO2 nanotubes through an high chemical dissolution rate of titanium and TiO2 in hydroflu-
AAO template-based titanium (IV) t-butoxide (Ti(OC4 H9 )4 ) hydrol- oric acid containing electrolytes and high H2 O content of the
ysis process. In their experiment, the wall thickness of the TiO2 electrolytes could result in limitation of length to few hundred
nanotubes can be controlled by adjusting the concentration of nanometers and non-ideal tubular shape of nanotubes [33,35].
Ti(OC4 H9 )4 solutions. Hence, the first generation of nanotubes length with approximately
In short, both types of template show the disadvantage of 500 nm or less could not meet the requirement of practical appli-
obtaining smaller nanotubes due to the restriction of pore sizes cations.
of the prepared template. Besides, it requires multi-step of pre- Subsequently, the use of buffered neutral electrolytes contain-
fabrication template and post-removal of template which are time ing various fluoride salts such as sodium fluoride [36] and ammonia
consuming for the practical applications. In the case of the post- fluoride [34] had been proposed to slow down the TiO2 dissolution
removal for hard template, array of nanotubes may be collapsed, rate. For this second generation of nanotubes, longer nanotubes
contaminated or become bundles onto the substrate during the with lengths up to several micrometers can be produced as the
wet-chemical etching in acidic or alkaline solution [30]. ultimate concentration of F− and hydrofluoric acid are determined
by the solution pH. Taking into account the importance of the pH
3.2. Electrochemical anodization method during anodization, Sreekantan et al. [34] reported that the reduc-
tion of acidity could decrease the chemical dissolution of titanium.
TiO2 nanotubes with ordered arrangement and high aspect ratio This allowed the formation of nanotube arrays up to 2.5 ␮m long.
can be produced by using electrochemical anodization method. On the other hand, electrochemical anodization in non-aqueous
Basically, TiO2 nanotubes grow on the surface of the anode of tita- organic solvent electrolyte containing F− can be considered as third
nium foils or thin sheets. The dimensions of nanotubes formed generation of nanotubes [37]. This method appeared as a promis-
in various electrolytes can be controlled by applying different ing approach to form smooth and long nanotube arrays. Various
electrolyte composition, applied voltage, pH and anodizing time types of organic solvents such as glycerol [37,38], ethylene glycol
[33,34]. The development of electrochemical anodization method [37,38], formamide [41], N-methylformamide [41] and dimethyl
132 Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142

sulfoxide [41] had been successfully employed for the formation electrical resistance of the cell [48]. As a result, more electric energy
of TiO2 nanotube arrays. Paulose et al. [41] found that the key to must be used and consequently increases the cost of operating
obtain very long nanotube arrays was to minimize the amount of electrochemical anodization.
water to be less than 5 wt.% during anodization process. Addition of
small amount of water allowed sufficient titanium oxidation rate 3.3. Hydrothermal method
since the oxidation process was difficult in pure organic electrolyte
[38,41]. Among these preparation methods, hydrothermal method has
According to the literature, the most widely accepted formation attracted more attention from the scientific community due to the
mechanism of TiO2 nanotubes during electrochemical anodization simple, cost-effective and environmental friendly method for large
involves several stages [35,38,42]. At the initial stage of anodiza- scale production of nanotubes. This synthesis conditions could be
tion, a dense TiO2 layer was formed on the titanium surface by adjusted to fabricate alternative low dimensional of TiO2 nano-
field-assisted oxidation via reaction (7). Ion migration occurred structures [10]. These nanostructures included TiO2 nanotubes,
within metal/metal oxide interface under the applied electric field. nanofibers, nanowires, nanoribbons and nanorods by controlling
Titanium ions could dissolve and react with oxygen ion (O2− ) or various parameters during hydrothermal synthetic conditions. The
hydroxyl ion (OH− ) provided by water to form TiO2 or titanium term nanotube was used for high aspect ratio materials with a
hydroxide, respectively. In the next stage, the titanium metal and hollow inner channel and nanowire for high aspect ratio materi-
oxide layer, which were in contact with the electrolyte, would etch als without an inner void. Meanwhile, nanofibers are similar to
by F− ions through chemical dissolution to form soluble hexaflu- nanowires but with significantly larger diameter [49]. Nanorods
orotitanium complex (Eq. (8)). The localized dissolution of oxide usually featured shorter aspect ratio as compared to nanowires.
resulted in the formation of tiny pits on the oxide layer and gradu- Fig. 3 shows the HRTEM images for the morphology of nanotubes,
ally grew into bigger pores with increasing anodization time. nanowires, nanofibers and nanorods, adapted from Yuan and Su
Ti + 2H2 O → TiO2 + 4H+ + 4e− (7) [50].
In year 1998, Kasuga et al. [51] first reported the synthesis of
+ − 2−
TiO2 + 4H + 6F → [TiF6 ] + 2H2 O (8) TiO2 nanotubes by employing an alkaline hydrothermal treatment
of TiO2 nanoparticles. With respect to the described methods, the
It is believed that the competition between rate of oxide growth alkaline hydrothermal route could produce nanotubes with tunable
at the metal/metal oxide interface (Eq. (7)) and the rate of oxide diameters (∼8–200 nm) [52]. After further investigation, major-
dissolution at the metal oxide/electrolyte interface (Eq. (8)) is the ity of the researchers found that the so-called “TiO2 nanotubes”
key factor in determining nanotube formation [35]. When the rate obtained via hydrothermal method were actually titanate nano-
of electrochemical etching equals to the rate of chemical dissolution tubes [53,54]. Thus, the term of titanate nanotubes was used for
of the nanotubes, the tube length remains the same. Thus, the final the subsequent review instead of TiO2 nanotubes.
stage determines the length of the nanotubes during the reaction. Although this synthesis method had been carried out exten-
The crucial key factors to achieve high degree arrangement of sively over the past decade, the crystalline structure, composition,
hexagonal close-packed nanotubes structured are anodic voltages thermal stability and formation mechanism are still in controversy.
for a given electrolyte before dielectric breakdown [43], purity of Several factors can influence the formation of titanate nanotubes
the material [43] and multi-step anodization [42]. By altering the such as phases and particle sizes of the starting materials [55], types
electric field during the growth of nanotubes, several types of tube and concentrations of the alkaline solution [56], hydrothermal tem-
geometry can be achieved such as bamboo- [44], nanolace- [45] and perature and duration [57] and post-treatment processes (acid
branched-type nanotubes [45]. For example, bamboo-type nano- washing and calcination) [58]. Table 4 shows several selected works
tubes were obtained by alternating voltage and holding time on on the conventional- [59], ultrasonication- [57,60] and microwave-
the microstructure [44], nanolace-type nanotubes were obtained hydrothermal [11,61] methods for synthesizing titanate nanotubes.
when extending the etching process of bamboo-type nanotubes The details of these combination techniques will be further dis-
[45], while tube branching morphology could be obtained when cussed in Section 4.4.
altering a higher to a lower anodization voltage [45]. Meanwhile, Ji The main disadvantage of this nanotube fabrication method
et al. [46] found that double-walled nanotubes with bamboo-like is the requirement of long reaction time in concentrated sodium
morphologies could be formed if the anodization process was car- hydroxide (NaOH) medium, which lead to the excessive intercala-
ried out under alternating voltage conditions. The double-walled tion of sodium ions in the produced nanotubes. Hence, more efforts
structure could be further tuned by heat treatment. have to be made in order to shorten the synthesize time and pro-
In addition to the dilute fluoride electrolyte approach, alter- duce ordered alignment of TiO2 nanotubes apart from the random
native approach using electrolyte containing perchlorate [14], alignment powder form [62].
chloride [39,47], bromide [47] or nitrate [36] ion is possible to
form bundles of TiO2 nanotubes on titanium substrate under very
rapid growth conditions. This method is known as rapid breakdown 4. Factors influencing the formation of titanate nanotubes
anodization and micrometers long nanotubes can be obtained in a via hydrothermal treatment
few min [39]. Hence, it is not easy to tune the morphology of tube
bundles into highly ordered self-organized TiO2 nanotubes through The microstructure of titanate nanotubes is easily affected by
this ultra-fast anodization technique. the starting materials, concentrations and types of the alkaline
Production of TiO2 nanotubes via electrochemical anodiza- solution, synthesize conditions (reaction time and synthesis tem-
tion method requires two/three electrode configurations with perature) and post-hydrothermal treatments (acid washing and
anodization voltage and the help of a power supply. Besides, the calcination temperature).
disadvantages of this method include the length distribution of
tubes is not well-defined [9], nanotubes are not separated from 4.1. Starting materials: phases and particle sizes
each other in an organized manner and do not have well-developed
gaps between the tubes over a large surface area [10]. Hydrogen The effects of different phases and particle sizes of the start-
gas is evolved at the cathode in the form of gas bubble nucleates, ing materials for titanate nanotubes are still debatable. Table 5
which will accumulate at the electrode surfaces and increase the summarizes several selected results on the formation morphology
Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142 133

Fig. 3. HRTEM images of (a) overall view nanotubes, detailed view on (b) nanotubes, (c) nanowires, (d) nanofibers and (e) nanorods.
Adapted from Yuan and Su [50].

of titanate nanotubes using different types of starting materi- both phase and particle size of the starting material. Pure anatase
als, alkaline solution and hydrothermal treatment. Morgan et al. phase with a particle size of 300 nm could produce nanotubes with
[55] observed that the dissolution process of anatase precursor 1 ␮m in length, micrometer sized of TiO2 with mixed phases pro-
corresponded to zero-order kinetic, where the dissolution was duced tubes with lengths between 200 and 500 nm, while P25
independent of precursor concentrations. Meanwhile, the disso- mixed phases with particle size of 30 nm produced lengths less than
lution kinetic for the rutile phase was second-order, indicated TiO2 100 nm. Viriya-empikul et al. [57] reported that interlayer spac-
rutile was affected by the diffusion of products and replenishment ing of nanotubes was not significantly affected by the size of the
of reactants to the surface. Anatase has often been observed to react starting material. They verified that smaller crystal size of start-
rapidly to produce nanotubes, whereas rutile has been observed to ing material could lead to titanate nanotubes with a higher surface
react incompletely with remaining unreacted rutile in the product area.
or nanorods shaped after similar hydrothermal conditions. They In addition, Morgado et al. [63] reported that TiO2 anatase raw
also observed that anatase with small crystalline and rutile phase materials with smaller crystal size usually led to faster rate of
with big crystalline size could produce nanotubes with lengths up thermal transformations of TiO2 nanotubes. Anatase with smaller
to several hundred nanometers, but the former starting material crystalline size could produce larger external diameter as com-
could produce nanotubes with smaller average external diameter. pared to those produced from anatase phase with bigger crystalline
Even though several authors reported that amorphous TiO2 size. However, Papa et al. [70] found that anatase particles with
could only produce nanofibers not nanotubes [50], Ylhäinen et al. a grain size of less than 17 nm could not form nanotubes at
[68] claimed to have obtained nanotube products from amorphous hydrothermal temperature of 150 ◦ C even with prolong treatment
precursors. Hence, it seems that there were other parameters play- duration from 36 h to 72 h. They concluded that too many small pri-
ing a role apart from crystalline form of TiO2 . Yuan and Su [50] mary germs involved in the dissolution/recrystallization reactions
reported that the effect of particle size of the starting material and it was hard to obtain super-saturation condition for the growth
toward the yield of the produced nanotubes was quite small in of the nanotubes. Meanwhile, Miao et al. [71] used 14 nm diame-
comparison to the hydrothermal temperature and concentration ter anatase aerogel powders as the starting material and produced
of alkaline solution. On the other hand, Dawson et al. [69] con- nanotubes after treatment duration of 40 h at 150 ◦ C. These con-
cluded that the length of titanate nanotubes was influenced by flicted results reported by both of the studies might probably due
134 Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142

Table 4
Hydrothermal method for preparation of titanate nanotubes.

Auxiliary methods Conditions hydrothermal reaction Dimensions of nanotubes References

Conventional hydrothermal • 5 g TiO2 powder in 10 M sodium hydroxide Average Do = 9 nm Kitano et al. [59]
(NaOH) Average Di = 6 nm
• 150 ◦ C for 20 h Interlayer spacing = 0.7 nm (2–4 walls)
• Acid washing with 0.1 M nitric acid (HNO3 ) L = 0.3–3.0 ␮m
Specific surface area = 423 m2 /g

Ultrasonication-assisted • 150 mg TiO2 powder in 30 mL of 10 M NaOH Average Do = 5 nm Zhu et al. [60]


hydrothermal • Mixture was sonicated at 280 W for 60 min Wall thickness = 1.3 nm
• 110 ◦ C for 4 h L = 200–300 nm
• Acid washing with 0.1 M HNO3
• 0.5 g TiO2 powder in 50 ml 10 M NaOH Di = 4–6 nm Viriya-empikul et al. [57]
• Mixture was sonicated at 7.6, 38.1 W for 8 min Interlayer spacing = 0.8 nm (2–6 walls)
• 150 ◦ C for 3 days Average hydrodynamic size of 53, 490, and
• Acid washing with 0.1 M hydrochloric acid 1760 nm at sonication power of 0, 7.6, and 38.1 W
(HCl) Specific surface area = 179, 258, and 248 m2 /g at
sonication power of 0, 7.6, and 38.1 W

Microwave-assisted • 0.5 g TiO2 rutile in 25 ml of 10 M NaOH Di = 3–5 nm Huang and Chien [11]
hydrothermal • Microwave power for 45 min at 200 ◦ C Do = 8–10 nm
• Acid washing with 0.1 M HNO3 L = 100–500 nm
Interlayer spacing = 1 nm (3–4 walls)
Specific surface area = 214 m2 /g
• 0.6 g P25 TiO2 in 70 ml 10 M NaOH Average Do = 9.9 nm Ou et al. [61]
• Microwave power of 70, 400, and 700 W at Average Di = 5.1 nm
130 ◦ C for 3 h Interlayer spacing = 0.82, 0.86, and 0.90 nm at
• Acid washing with 0.5 M HCl microwave power of 70, 400, and 700 W (4–5 walls)
Specific surface area = 323, 367, and 320 m2 /g at
microwave power of 70, 400, and 700 W

to the different conditions used such as fabrication vessels, amount was observed in 10 M lithium hydroxide (LiOH) at the temperatures
of starting materials and concentrated alkaline solution. of 120 and 150 ◦ C [56], whereas flower-like structure on the TiO2
surface could be obtained when treated in 10 M LiOH solution at
temperature of 80 ◦ C [75].
4.2. Alkaline solution: types and concentrations

The alkaline type and their concentrations play important role


in hydrothermal process. The most favorable titanate nanotubes 4.3. Hydrothermal treatment: temperature and duration
were prepared via hydrothermal treatment of TiO2 powders in a
10 M NaOH solution [63,70]. There was only small proportion of In general, the amount, length and crystallinity of titanate nano-
nanotubes formed when NaOH concentration was lower than 5 M tubes increase with the hydrothermal temperature [76]. However,
or higher than 18 M [56]. The concentration of dissolved Ti4+ in solu- there exists an optimal temperature, which usually reported in
tion can determine the rate of crystallization of titanate nanosheets, the range of 130–150 ◦ C to obtain high purity of nanotubes with
which in turn control the morphology of the final nanostructures high surface area [49,57]. During higher reaction temperatures
[10]. (>180 ◦ C), the main products were nanoribbons or nanowires like
Bavykin et al. [72] found that an increment in the molar ratio morphologies [19]. Yuan and Su [50] reported that TiO2 nanotubes
of TiO2 to NaOH generally resulted in higher average pore diam- could be formed in the reaction temperature ranged from 100
eter and a reduction in the specific surface area. Meanwhile, the to 180 ◦ C when either crystalline anatase or rutile or commercial
pore volume of nanotubes was independent of the molar ratio of P25 were used as the raw materials. Nanoribbons were produced
TiO2 to NaOH. Furthermore, Huang et al. [64] demonstrated that when increasing reaction temperature from 180 to 250 ◦ C, while
nanosheets, nanotubes and nanowires were three kinetic products nanowires were obtained when reaction temperature ranged from
from the reaction of TiO2 with NaOH when increasing alkaline con- 130 to 240 ◦ C if KOH were used as the alkaline solution.
dition from 5 to 12 M. They observed that the nanowires had wider Bavykin et al. [72] found that increasing temperature from 120
diameter and longer length as compared to nanotubes morphol- to 150 ◦ C for 22 h resulted in larger average diameter of titanate
ogy. However, Sikhwivhilu et al. [56] reported that a small portion nanotubes. Subsequent increment of temperature resulted in a
of shorter nanotubes was produced when concentration of NaOH sharp reduction in pore volume of produced samples and wider
solution increased to 18 M. Thus, the concentration of NaOH could size distribution in the diameter of fibers morphology. However,
influence in the yield of tubular morphology. Sreekantan and Wei [76] observed that the average outer diame-
Furthermore, Bavykin et al. [73] discovered that titanate nano- ters of the nanotubes produced at 110, 130 and 150 ◦ C for 24 h were
tubes could be prepared using 10 M potassium hydroxide (KOH) similar, at about 10 nm. Seo et al. [66] observed that increasing the
and low temperature of 56 ◦ C after 12 days. Thus, titanate nano- synthesis temperature up to 150 ◦ C would improve crystallinity of
tubes could also be produced using a NaOH/KOH binary aqueous the tube and increased tube lengths from few hundred nanometers
mixture at 100 ◦ C [74]. Recently, Sikhwivhilu et al. [56] found to few micrometers, but with the same diameters. Meanwhile, Lan
that potassium titanate nanotubes could also be produced by et al. [67] found that the outer diameter of the nanotubes increased
hydrothermal treatment in 10 M KOH at 120 and 150 ◦ C for slightly when increasing hydrothermal temperature above 150 ◦ C.
24 h. They reported that KOH was more reactive with TiO2 to This indicates that reaction temperature is a key factor in deter-
yield different shapes of nanostructured materials compared to mining the overall morphology of the product. Hence, the effect
NaOH. Ammonium hydroxide solution showed no influence on the of hydrothermal temperature on the formation of the nanotubes
microstructure of TiO2 . Meanwhile, similar spherical morphology products is similar to that of the alkaline concentration.
Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142 135

Table 5
Effects of starting materials, alkaline solution and hydrothermal treatment on the morphology of titanate nanotubes.

Starting materials (phase, Alkaline solution Hydrothermal treatment Dimensions of nanotubes References
composition, particle size) (types, concentration) (temperature, duration)

(i) Anatase, >95%, 112 ± 1 nm 9 M NaOH (i) 160 ◦ C, 2–8 h (i) L = several hundred nanometers, average Morgan et al. [55]
(ii) Rutile, >93%, 320 ± 10 nm (ii) 160 ◦ C, 4 h Di = 3.7 ± 0.7 nm, average Do = 9.4 ± 1.7 nm,
interlayer spacing = 0.74–0.75 nm (3–5 walls)
(ii) L = several hundred nanometers, average
Di = 3.6 ± 0.7 nm, average Do = 9.8 ± 1.5 nm,
interlayer spacing = n.d. (6–7 walls)

(i) Anatase, 220 nm 10 M NaOH (i) 120 ◦ C, 30 h (i) L = several hundred nanometers, average Morgado et al. [63]
(ii) Anatase, 8 nm (ii) 120 ◦ C, 15 h Di = 3–5 nm, average Do = 6–10 nm, interlayer
spacing = n.d. (4–5 walls), pore size
distribution = 2–5 nm; 5–60 nm
(ii) L = several hundred nanometers, average
Di = 4–10 nm, average Do = 10–30 nm, interlayer
spacing = 0.74–0.75 nm (>5 walls), pore size
distribution = 2–10 nm

Anatase (i) 3 M NaOH 180 ◦ C, 24 h (i) Nanosheets Huang et al. [64]


(ii) 5 M NaOH (ii) Nanotubes, L = several hundred nanometers,
(iii) 10 M NaOH average Di = 2–4 nm, average Do = 7–10 nm,
(iv) 12 M NaOH interlayer spacing = n.d. (3–5 walls)
(v) 15 M NaOH (iii) Mixture of nanotubes and nanowires,
L = several tens of micrometers, average
D = 40–250 nm
(iv) Mixture of nanowires and amorphous particles

P25 Degussa (i) 5 M NaOH 140 ◦ C, 24 h (i) Fibrous structures, L = 100–200 nm, average Camposeco et al. [65]
(ii) 7 M NaOH D = 25–140 nm
(iii) 10 M NaOH (ii) Nanotubes structures, average Di = 3–4 nm,
(iv) 12 M NaOH average Do = 8–10 nm, wall thickness = 2.5 nm
(iii) Nanotubes structures, average Di = 4 nm,
average Do = 8 nm, wall thickness = 3.5 nm,
interlayer spacing = 0.7 nm (2–4 walls)
(iv) Nanowires structures, L = 300–500 nm, average
D = 15–20 nm

P25 Degussa, 25 nm 10 M NaOH (i) 70 ◦ C, 48 h (i) 2D nanosheets Seo et al. [66]


(ii) 90 ◦ C, 48 h (ii) Nanosheet and fiber-like mixed phase
(iiii) 110 ◦ C, 48 h (iii) 1D fiber/tube-like, L = few hundred
(iv) 150 ◦ C, 48 h nanometers, average Do = 10–20 nm
(iv) 1D separated fibers/tube, L = few micrometers,
average Do = 10–20 nm

Rutile powder, 120–280 nm 10 M NaOH (i) 100 ◦ C, 48 h (i) Nanotubes, average Do = 7–11 nm Lan et al. [67]
(ii) 125 ◦ C, 48 h (ii) Nanotubes, average Do = 8–12 nm
(iii) 150 ◦ C, 48 h (iii) Nanotubes, L = 50–200 nm, average
(iv) 180 ◦ C, 48 h Di = 2–3 nm, average Do = 8–15 nm
(iv) Nanorods, L = several micrometers, average
D = 40–120 nm

Note: n.d. is defined as no data provided.

At relatively low reaction temperature around 90 ◦ C, nanosheets influence on the morphology and properties of titanate nanotubes
[76], nanofibers [66] or nanotubes [57] could be formed after 24 h, [57], apart from dispersing TiO2 uniformly in alkaline solution.
48 h, 72 h, respectively. This proved that the yield of nanotubes Meanwhile, microwave irradiation is able to deliver energy to
could increase with duration treatment. However, prolonged reac- the reactants through molecular interactions and heat treatment
tion duration (72 h to 1 week) might lead to the transformation using microwave is an economic, rapid and homogeneous heating
of nanotubes to nanofibers or nanowires [77]. The width of the method [11].
obtained nanofibers or nanowires was about 100–300 nm, which Zhu et al. [60] found that titanate whiskers and nanotubes
was significantly larger than the diameter of nanotubes. Thus, an were obtained when sonication powers used were 580 W and
increment in the average nanotubes diameter can be observed 280 W, respectively and the synthesis duration was shortened from
at longer synthesis time. Hence, the treatment duration could be 20 to 4 h. Meanwhile, nanosheet-, nanofiber- and nanotube-like
reduced if higher reaction temperature was performed in order to morphologies were observed under sonication of 100, 280 and
obtain nanotubular structures. 380 W, respectively [78]. Appropriate sonication power and dura-
tion may help the reaction of sodium and hydroxyl species on
4.4. Assisted hydrothermal synthesis the de-agglomerated TiO2 particles and the dissolution/breaking
of Ti O Ti bonding. This resulted in smooth unhindered peeling
The purposes of the application of sonication- or microwave- of large nanosheets and consequently rolled into longer titanate
assisted hydrothermal treatment are due to the advantages such nanotubes [57,78]. Ou et al. [61] reported that higher microwave
as shorter synthesis duration [11,60] and enhance the yield irradiation power was beneficial to the intercalation of Na atoms
and length of titanate nanotubes [57]. The power of sonication into titanate nanotubes in order to enlarge the spacing between
and irradiation duration played an important role in fabricat- interlayers. Titanate nanotubes could be obtained after 45 min
ing titanate nanostructured materials. Sonication showed a strong under microwave irradiation [11].
136 Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142

Fig. 4. Schematic representation of the thermal transformations pathways as a function of sodium content in TiO2 nanotubes [53].

In addition, organic solvent-controlled hydrothermal synthe- Meanwhile, several researchers suggested that acid washing
sis had been investigated to achieve different 1D nanostructures was only applied for the ion-exchange process as they found that
morphology of TiO2 . Solvents with different physical and chemi- titanate nanotubes were formed during hydrothermal reaction
cal properties have a strong effect on the solubility, reactivity and [53,55,66]. When proton replaces sodium ion in the structure, acid
diffusion behavior of the reactants, which are able to influence the washed titanate nanotubes possesses stronger hydroxyl bonds,
structure and morphological features [79]. It had been reported that results in decrement of interlayer spacing and increment in specific
the samples prepared using ethanol [80], methanol [80], methyl surface area [53]. Papa et al. [70] found that acid washing turned
ethyl ketone [80] and ethylene glycol [81] exhibited nanotube out to be necessary when nanoribbons structure was obtained after
structures. However, Das et al. [81] found that nanosheets were hydrothermal treatment. Researchers often reported an optimum
transformed to nanorods and nanowires in water–ethanol and concentration of hydrochloric acid (HCl) during acid washing in
water–ethylene glycol mixed solvents, respectively. Meanwhile, order to prepare sodium-free nanotubular structure effectively.
Santara and Giri [82] obtained nanoporous nanoribbons products in Further increment of the concentration of HCl would destroy nan-
water–ethylene glycol mixed solvents. On the other hand, titanate otubular structures [58].
nanotubes with ultrahigh crystallization can be obtained after The quantity of sodium ions plays a crucial role in the mor-
hydrogen peroxide treatment under reflux at 40 ◦ C for 4 h [83]. This phological, structural stability and phase transformation of the
proves that titanate microstructures are more easily affected by the nanotubes [53]. The detailed discussions on the role of sodium
preparation conditions, rather than the types of organic solvent. in stabilizing the nanotubular structure of titanate nanotubes
had been reported by Zhang et al. [85]. They demonstrated that
4.5. Post-hydrothermal treatments heat-treatment of sodium titanate nanotubes may induce the dehy-
dration of interlayered OH groups, whereas hydrogen titanate
In order to achieve high catalytic activity of titanate nanotubes nanotubes experienced dehydration of intralayered and interlay-
and control the tubular structure of nanotubes, the investigation of ered OH groups. These types of dehydrations did not only reduce
post-hydrothermal treatment has received more attention com- the interlayer distance and nanotube length, but also destroyed the
pared to hydrothermal synthetic conditions. Post-hydrothermal nanotube morphology.
treatments include acid washing and calcination steps. Acid wash-
ing might affect the nanotubes formation [84], crystalline structure
or even chemical composition of the final nanotubes product [53]. 4.5.2. Calcination
Meanwhile, it is believed that calcination process can induce the The level of sodium ions in titanate nanotubes had been iden-
transformation of titanate to anatase phase, modify microstructure tified as an important factor that can influence thermal stability
based on the amount of sodium present in nanotubes, improve crys- and phase transformation of the materials [53]. Hence, the ther-
tallinity and reduce surface defects of the material [79]. From the mal behavior of titanate nanotubes could be categorized into three
point of view of practical usage, it is an important issue to study the groups: high, medium and low or non-sodium titanate nanotubes.
structural stability of the titanate nanotubes and their correspond- The structural changes in high, medium and low sodium titanate
ing microstructure changes at various calcination temperatures. nanotubes generally follow the sequence as shown in Fig. 4 [53].
High sodium titanate nanotubes without acid washing showed
4.5.1. Acid washing the best thermal stability and it could remain its nanotubu-
Kasuga et al. [84] considered that the acid washing process was a lar structure when increasing temperature from 300 to 500 ◦ C
crucial step for the formation of titanate nanotubes. They proposed [85]. The coexistence of sodium hexatitanate (Ti6 O13 ) and rutile
that the purpose of acid washing was to remove electrostatic repul- appeared after calcined at 600–800 ◦ C. A few studies found that
sion of residual charge, in which Ti O Na converted to Ti OH. the tubular structures of medium sodium titanate nanotubes can
Then, the dehydration of Ti-OH bonds could result in the formation be maintained up to 300–400 ◦ C [53]. Conversion of titanate nano-
of nanotubes from nanosheets. tubes into anatase occurred at about 500 ◦ C. Meanwhile, low or
Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142 137

Fig. 5. The structure models of (a) anatase TiO2 [88], (b) trititanate H2 Ti3 O7 [87],
(c) dititanate H2 Ti2 O5 ·H2 O [88] and (d) lepidocrocite-type titanate (Hx Ti2−x/4 Vx/4 O4 ,
where x ∼ 0.7, V represents vacancy) [87]. Fig. 6. Schemes of three possible mechanisms for multi-walled titanate nanotubes
formation, include (a) the helical scrolling of a single-layer nanosheet, (b) the curv-
ing of several conjoined nanosheets and (c) direct production of a multi-layered
non-sodium titanate nanotubes could only retained the tubu-
nanotube [72].
lar shape at 250–350 ◦ C. Further increment of temperature, they
transformed into the intermediate phases of hydrogen Ti6 O13 , fol-
lowed by monoclinic and metastable TiO2 (B) [58]. Anatase usually by every nth octahedral [77]. Lepidocrocite-type titanate lattice
appeared during calcination temperature above 450 ◦ C. This phase is composed of two-dimensional (2D) lepidocrocite-type or flat
transformation temperature might be different at varying synthetic host layers sheets in which TiO6 octahedra are connected to each
conditions. other via edge-sharing [91]. This depicts that the preparation
In addition, annealing as-prepared titanate nanotubes at ele- conditions play an important role in determining morphologies
vated temperature in an oxygen-poor environment such as a and microstructures of titanate nanostructures.
pure helium, nitrogen, argon gas atmosphere or vacuum condi- The mechanism of formation of the various types of nanotubes
tion would create oxygen vacancies [86]. Zhang et al. [85] reported has been the subject of intense research. According to the liter-
that single electron trapped oxygen vacancy and hydrogen vacancy ature, acid washing after hydrothermal was essential for sodium
were produced during intralayered dehydration of titanate nano- exchange purpose as well as formation of nanotubes [51,84]. Mean-
tubes. However, these as-formed oxygen vacancies on titanate while, strong evidences state that the key step for the formation
nanotubes will slowly disappeared while the material was exposed of nanotubes is during hydrothermal reaction [53]. However, it
to air, even at room temperature [86]. Hence, modification of is widely accepted that the formation mechanism of nanotubes
titanate nanotubes can be considered to stabilize the oxygen vacan- follows the 3D → 2D → 1D model [92]. Chemical bonds in the start-
cies. ing 3D TiO2 will break under hydrothermal reaction, in which 2D
layered entities will be formed and then convert into 1D nano-
5. Crystal structure, formation mechanism and tubes through sheet folding mechanism. Bavykin’s group [72] had
structural-activity of titanate nanotubes proposed three possible mechanisms for the formation of multi-
walled titanate nanotubes. These included (a) the helical scrolling
A number of crystalline structures for as-synthesized titanate of a single-layer nanosheet, (b) the curving of several conjoined
nanotube materials including anatase TiO2 [84], lepidocrocite- nanosheets and (c) direct production of a multi-layered nanotube,
type titanate (Hx Ti2−x/4 Vx/4 O4 , where x ∼ 0.7, V represents as shown in Fig. 6.
vacancy) [87], dititanate acid (Nax H2−x Ti2 O5 ·H2 O, where 0 < x < 2) Among these mechanisms of formation, the wrapping of a sin-
[54,88], trititanate acid (Nax H2-x Ti3 O7 ) [53] and tetratitanate gle nanosheet has been favored by a number of research groups
acid (H2 Ti4 O9 ·H2 O) [89] had been proposed. Fig. 5 shows the [53,93]. This would result in the inequality number of wall layers
structure models of anatase TiO2 [88], H2 Ti3 O7 [87], H2 Ti2 O5 ·H2 O on two sides of nanotubes and interlayer spacing of approximately
[88] and lepidocrocite-type titanate (Hx Ti2−x/4 Vx/4 O4 , where 0.78 nm. Zhang et al. [93] indicated that the imbalance charge of
x ∼ 0.7, V represents vacancy) [87]. The crystal system of anatase hydrogen or sodium ion concentration at two different sides of
is tetragonal, lepidocrocite-type titanate and H2 Ti2 O4 (OH)2 are nanosheets may introduce excess surface energy to strain driv-
orthorhombic, while H2 Ti3 O7 and H2 Ti4 O9 ·H2 O are monoclinic ing the peeling process and resulting in wrapping of nanotubes.
structures. Similar to anatase TiO2 , all titanate acids consist of This kind of hydrogen deficiency driving force in surface layers is
edge-shared TiO6 octahedra and zigzag ribbons configurations unable to bend multiple layers of nanosheets although it can be
[90]. The arrangement of layers of TiO6 octahedrons is diverse in done through mechanical stress. On the other hand, Kukovecz et al.
different crystal structures. Three edge-sharing octahedral is in [94] argued that particle surfaces were cut up into small terraces
the unit cell of trititanate acid while four edge-sharing octahedra before curving up of the nanoloop resulted from nanosheets.
is in tetratitanate acid. Particularly, titanate acid Tin O2n+1 (where Meanwhile, Bavykin et al. [72] demonstrated the imbalance
n = 3–5) consists of parallel corrugated layers of edge-sharing width of different layers during crystallization and rapid growth
TiO6 octahedra, where octahedra joins at the corners and stepped of layers which might lead to nanosheet with a single curved layer
138 Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142

and then several conjoined curved layer in order to decrease the temperature. In contrast, Turki et al. [58] observed that photo-
excess energy. They suggested that formation of nanotubes under- catalytic activity of free sodium titanate nanotubes calcined at
went several steps: (i) dissolution of TiO2 accompanied by epoxial 400 ◦ C showed 4 times more active than P25 for the degradation
growth of layered nanosheets; (ii) dissolution–recrystallization of of formic acid. Meanwhile, Lee et al. [98] found that medium con-
nanosheets; (iii) curving of nanosheets and (iv) exchange of sodium tent of sodium titanate nanotubes possessed higher photocatalytic
ions to proton during washing of nanotubes. activity as compared to low and high medium content of titanate
Morgado et al. [95] found that small crystal TiO2 possessed fast nanotubes. However, its activity was still lower than P25 in the pho-
dissolution rate which accelerated recrystallization of nanosheets. tocatalytic oxidation of a basic dye. All these results indicated that
When the rate of nanosheet crystallization was higher than the rate the role of sodium was not precisely evaluated while the relation-
of their curvature into nanotubes, the formation of thicker planar ship between photocatalytic activity and morphology of titanate
entities with multiple layers would be favored before they could nanotubes was not clearly defined.
be easily rolled up. Thus, small crystal TiO2 led to low yields of
nanotubes and non-hollowed layered particles with thicker walls 6. Modification of titanate nanotubes as photocatalysts
and larger external diameter. They also suggested that this was the
reason for those researchers who were not able to obtain nanotubu- It should be noted that the photocatalytic activity of TiO2 is
lar materials from freshly precipitated amorphous TiO2 due to the limited by the light absorption properties, electron–hole recom-
fast dissolution rate under alkali solution. bination rate, reduction and oxidation reaction rates on the surface
Both titanate and anatase structures contain zigzag ribbons of by the electrons and holes [1,79]. Many efforts on the modifica-
TiO6 octahedra that share four edges with each other [96], while tion of titanate nanotubes have been made to modify the electronic
rutile structure is composed of TiO6 octahedrons that share two band structure in order to promote the separation of electron–hole
edges with each other [97]. This can be surmised that anatase pair and improve catalytic activity. Titanate nanotubes can be mod-
easily converted into titanate nanotubes, while rutile requires a ified by various strategies such as deposition of metal nanoparticles,
higher temperature in order to convert into titanate nanotubes single- and co-doping with two or more foreign ions, coupling
due to the structural dissimilarity. Dawson et al. [69] reported with other semiconductors and hybrid with carbon nanomateri-
that pure anatase phase with large particle size able to allow the als. Several recent research works on the modification of titanate
dissolution–recrystallization processes and no other phase impu- nanotubes classified according to different methods have been
rity to impede greater growth for the nanotubes. The mixture of summarized in Table 6.
phases in the starting material would inhibit the reaction of anatase
and thus produced shorter tubes than the pure phase. The rutile
structure cannot simply undergo the dissolution and exfoliation 6.1. Metal-loaded
processes, thereby, total structural destruction and recrystalliza-
tion become necessary [97]. Metal loading on titanate nanotubes surface is one of the
Zhu et al. [96] revealed that the phase transition from titanate methods used to improve the catalytic activity. Various metal
to anatase was a topochemical reaction process, rather than dis- nanoparticles such as Pt [100], Au [100] and Ag [1] have
solution of titanate and recrystallization of anatase. They found been employed in titanate nanotubes through photo-deposition,
that the anatase retained the morphology of the reactant titanate deposition–precipitation or ion-exchange. It is well known that
because the large structural units remained relatively unchanged Schottky junction formed by the interface of metal–semiconductor
during dehydration process. It is noted that more investigations, [100]. The noble-metal particles have the property of collective free
evidences and discussions need to be done in order to support the electron charge at the interface when particles excited by light,
formation mechanism of titanate nanotubes. known as the process of surface plasmon resonance [1]. When these
In general, photocatalytic activity of titanate nanotubes was noble-metal particles load on to titanate nanotubes surface, the
determined by three structural factors which were specific surface plasmon-induced electrons are transported from the photo-excited
area, pore volume and amount of anatase phase in titanate nano- noble-metal particles to the conduction band of the titanate nano-
tubes [98]. It was often reported that larger specific surface area tubes [100]. This facilitates electron–hole separation, promotes the
would allow more reactants to adsorb to the surface of the cata- interfacial electron transfer process and enhances photocatalytic
lyst, which was beneficial for the subsequent oxidation reaction. performance of this hybrid catalyst [1,2]. Besides, metal ions with a
Higher pore volume would result in more rapid diffusion of var- standard redox potential is more positive than the conduction band
ious compounds during the photocatalytic reaction. Moreover, it of TiO2 can be photo-reduced to the low-valent states and then be
is well known that more anatase phase can induce faster photo- deposited on the surface of TiO2 nanomaterials [15].
catalytic reaction. Turki et al. [58] demonstrated that large surface
area was not the only crucial factor to achieve the highest pho- 6.2. Doping
tocatalytic activity. A good crystalline quality of anatase phase and
coherent crystallographic domains of TiO2 must be fulfilled in order According to previous research, metal or non-metal ion doping
to achieve high photo-degradation activity. tends to narrow the band gap energy, increases separation of
A straight forward comparison of the photocatalytic activity electron–hole pairs and inhibits recombination photo-generated
results obtained in different studies is difficult because final prod- electron–hole pair [79,112]. Thereby, the overall catalytic activ-
ucts with different specific structural and optical properties can ity of the catalyst can be enhanced. Various methods have been
be obtained. For example, in the literature, the amount of sodium introduced to prepare doped titanate nanotubes through chemical
content present in the titanate nanotubes might lead to different (sol–gel, co-hydrothermal) or physical method (physical mixing).
photocatalytic activities. According to Qamar et al. [99], titanate It had been reported that the dopant cation with almost similar
nanotubes with high, medium or free sodium content without ionic radius was easier to substitute Ti4+ cation and more difficult
calcination showed no photocatalytic activity for the degradation to replace an oxygen ion with other anion due to the difference
of the amaranth dye. The photocatalytic activities of free sodium in charge states and ionic radius [79]. In principle, the formation
titanate nanotubes calcined at 400 ◦ C and medium sodium titanate of oxygen vacancies on titanate nanotubes during doping process
nanotubes calcined at 500 ◦ C were less active than P25 though leads to the creation of unpaired electrons or Ti3+ centers in order
they reached their maximum activity with increasing calcination to keep the charge balance [86]. The excess electrons located on the
Table 6
Several recent studies on modified-titanate nanotubes and their effects.

Modifier Preparation methods, modifier source Modification effect (i) Application Reference
(ii) Important findings

Metal deposition
Copper (Cu) Hydrothermal, photodeposition Cu served as the electron mediator to prolong (i) Photocatalytic degradation of bisphenol A (BPA) Doong et al.
Source: 0.5–2.0 wt.% copper (II) nitrate the retention time of photo-generated radicals (ii) The pseudo first-order rate constants for BPA [15]
pentahemihydrate photo-degradation by 1 wt.% Cu-deposited TiO2 /TNTs
were 1.8–5.2 and 4.3–12.7 times higher than those of
pure Degussa P25 and anatase TiO2
Gold (Au) and platinum (Pt) Hydrothermal, photodeposition Au or Pt formed a higher potential gradient of (i) Photocatalytic degradation of methyl orange Zhao et al.
Source: 1 wt.% Schottky junction than at the TiO2 /electrolyte (ii) The degradation rates of methyl orange solution [100]
chloro-platinic acid (H2 PtCl6 ·6H2 O) and auric interface, which facilitated charge separation were 96.1%, 95.1%, 88.9% and 82.6% in the presence of
chloride acid Au–TNTs, Pt–TNTs, TNTs and P25 catalysts,
respectively after 40 min
Silver (Ag) Hydrothermal, deposition–precipitation Ag nanoparticles acted as additional separator (i) Photocatalytic degradation of caffeine Plodinec et al.
Source: 1 at.% silver nitrate (AgNO3 ) of the electron and hole pair (ii) The hydrogenated TNTs@Ag sample showed two [1]
times photocatalytic activity higher than the

Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142


hydrogenated TNTs
Ag Hydrothermal, deposition–precipitation and Ag–AgCl–TNTs hybrid enhanced visible light (i) Photocatalytic degradation of methylene blue Tang et al. [2]
AgX photo-deposition absorption which induced by the surface (ii) Ag–AgCl–TNTs was found to possess the highest
Ag–AgX (X = Cl, Br, I) Source: AgNO3 , hydrogen halide plasmonic effect methylene blue removal rate as compared to Ag–AgCl
and Ag–AgBr
Single- and co-doping
Carbon (C) and C–tungsten (W) Sol–gel, hydrothermal W acted as an electron and hole recombination (i) Photocatalytic degradation of 4-chlorophenol Neville et al.
Source: 1 at.% melamine borate (as C dopant) center. C-doped material exhibited high (ii) The degradation rates of 4-chlorophenol were [18]
and tungstic acid (C and W dopants) surface area, enhanced optical properties and 36.1%, 31.4%, 24.3% and 10% in the presence of C–TNTs,
charge separation C–W–TNTs, TNTs and P25, respectively after 1 h
Silver ion (Ag3+ ) Hydrothermal, photo-reduction Ag nanoparticles acted as electron acceptor (i) Photocatalytic degradation of ceftiofur sodium (CFS) Pugazhenthiran
Source: 2.5 wt.% AgNO3 and transferred to the adsorbed O2 effectively (ii) Photocatalytic degradation of CFS was about 6 et al. [3]
times higher than bare TNTs
Aluminum ion (Al3+ )–iron ion (Fe3+ ) Co-hydrothermal Fe3+ acted as shallow trapping sites. Al–Fe (i) Photocatalytic degradation of humic acids Yuan et al.
Source: 0.1–5.0 at.% aluminum chloride and co-doping narrowed band gap energy and (ii) Al–Fe co-doping at 0.25: 0.75 at.% showed the [101]
iron(III) nitrate promoted the generation of OH radicals highest degradation efficiency 79.4% as compared to
1 at.% Fe-doped (77.4%) and 1 at.% Al-doped (60.4%)
Pt–nitrogen (N) Hydrothermal, photo-deposition, Pt doping used as electron trapper to prevent (i) Photocatalytic production of hydrogen from Slamet et al.
Source: 0.6 wt.% H2 PtCl6 and 0.5 wt.% charge recombination and increased stability glycerol–water mixture [102]
ammonium hydroxide of N in catalyst (ii) Pt–N–TNTs produce H2 amounts approximately 13
times, 6 times, and 20% higher as compared with P25,
TNTs, and Pt–TNTs, respectively
N Co-hydrothermal N doping narrowed the band gap of TiO2 and (i) Photocatalytic degradation of rhodamine B Hu et al. [103]
Source: Weight ratio of TiO2 :urea was 1:1–1:10 served as electron trapper to inhibit the charge (ii) N-doped TNTs at weight ratio of 1:2 exhibited the
recombination highest photocatalytic activity than the undoped TNTs
and P25
Gadolinium ion (Gd3+ )–N Hydrothermal, ion-exchange Gd3+ acted as electron-trapped agent to inhibit (i) Photocatalytic degradation of rhodamine B Liu et al. [104]
Source: Molar ratio of gadolinium (III) the charge recombination. Co-doping inhibited (ii) The degradation rates were 81.4%, 61.5% and 37.1%
nitrate:ammonium chloride were 1:1, 1:2 and the particle growth and resulted in higher in the presence of 1:1 Gd–N–TNTs, Gd–TNTs and
2:1 crystallinity N–TNTs, respectively after 2 h
C–N–sulfur (S) Hydrothermal, physical mixing, calcination C acted as a photo-sensitizer to inject electron (i) Photocatalytic degradation of toluene Dong et al.
Source: Molar ratio of Ti:thiourea were 1:1, 1:2 to TNTs. N–S co-doping modified the electronic (ii) The molar ratio of Ti:triourea of 1:1 and 1:2 showed [105]
and 1:3 band structure and narrowed the band gap the highest pseudo first-order reaction rate constants
energy of TNTs under UV–vis light and visible light, respectively
Coupled semiconductors/binary
composites
ZnO Hydrothermal, wet-impregnation ZnO–TNTs facilitated charge separation, (i) Photocatalytic degradation of rhodamine B Wang et al.
Source: 10–30 wt.% zinc acetate increased the lifetime of the charge carriers (ii) 20 wt.% of ZnO–TNTs nanocomposites exhibited the [106]
and interfacial charge transferred to adsorbed highest photocatalytic activity than pure TNTs, P25
substrates and ZnO

139
140 Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142

oxygen vacancy states could serve as the centers to capture elec-

Kim et al. [109]

Qianqian et al.
trons [86]. This facilitates the efficient adsorption of oxygen and

Li et al. [108]
Grandcolas
et al. [107]
water on titanate surface and reaction with electrons and holes,

Jiang et al.
Reference

leading to the formation of O2 − and OH− , respectively. However,

[111]
[110]
these oxygen vacancies can also act as recombination sites in the
presence of high amount of dopants [112]. Hence, an optimum
doping amount (with an acceptable amount of oxygen vacancies)

(ii) TiO2 –TNTs composites on the CNFs surface showed


within 4 h), this followed by 10 wt.% CdS–TNT and TNT

photocatalytic activity (more than 90% within 40 min)


could exhibit better visible light absorption and higher photocat-
(i) Photocatalytic degradation of diethylsulfide (DES)

highest degradation efficiency (94%) as compared to


(ii) 10 wt.% PbS–TNT showed the best activity (95%

the highest photocatalytic activity as compared to


(i) Photocatalytic degradation of gaseous CH3 CHO

(ii) CNTs–TNTs at a mass ratio of 0.10 showed the


alytic activity.
efficiency and resistance to sulfate deactivation
(i) Photocatalytic degradation of methyl orange

(i) Photocatalytic degradation of methyl orange


photocatalytic activity in terms of DES removal
(ii) 4 wt.% WO3 –TNTs composite enhanced the

Titanate nanotubes doped with metal ions such as transition

(i) Photocatalytic degradation of rhodamine B


pure TNTs (86%) and CNTs (21%) after 40 min

(ii) 5 wt.% graphene–TNTs possess the best


metals (Al [101] and Fe [101]), rare earth metals (Gd [104]) and
noble metals (Ag [3]) have been widely investigated to enhance
the photocatalytic activity of the titanate nanotubes. Among these
elements, Fe3+ has been found to be the most used dopant due to its
low band gap energy of 2.6 eV and the radius of Fe3+ (0.64 Å) is quite
TNTs–CNFs and TiO2 –CNFs

similar to that of Ti4+ (0.68 Å) [113]. A new energy level within the
(ii) Important findings

band gap of TiO2 would be introduced when Fe3+ ions substituted


Ti4+ sites, which enabled visible light absorption. This is attributed
to the charge transfer transition between the d-state electron of the
(i) Application

dopant and the conduction band of TiO2 or from valence band to


the new energy level [114].
Moreover, Fe3+ dopants could act as hole and electron trapped
sites and change the oxidation states to Fe4+ and Fe2+ , respectively
[113,114]. According to crystal field theory, the electronic config-
reduced charge recombination. Adsorptivity of

urations of Fe4+ and Fe2+ ions are relatively unstable as compared


light, facilitated the charge photo-generated

PbS and CdS acted as the sensitizers to light

The UV light absorption and surface area of

to Fe3+ ions (half-filled high spin d5) [114]. The electrons and holes
WO3 –TNTs increased absorption of visible

CNTs were good electronic conductor and

organic pollutant on CNTs was increased


absorption, large BET surface areas, and

are easily de-trapped to the adsorbed oxygen and surface hydroxyl


groups, respectively and restores to its half-filled electronic con-
graphene–TNTs and generated the
separation and prevented charge

Graphene acted as a sensitizer in

figuration. Therefore, the recombination rate of electron and hole


pairs can be retarded with the introduction of Fe3+ . Besides, the
increased separation charge

CNFs–TNTs were increased

photo-generated electrons

effect of Fe doping is highly dependable on the dopant amount as


the doping sites could also serve as charge recombination centers
Modification effect

when present in high concentration [113]. The additional benefit of


Fe doping is that the trace amount of Fe that might be leaching out
recombination

from the photocatalyst could undergo fenton-like reactions [115].


Cationic doping materials suffer from thermal instability
Carriers

and cationic doping could be localized in the d-state deep of


titanate band gap, which usually act as recombination center for
electron–hole pairs [103]. Hence, non-metal ion doping appears
Source: 1, 5 and 10 wt.% commercially available
Microwave-assisted hydrothermal, dip-coating

polyacrylonitrile solutions via electrospinning

as a promising way to avoid the disadvantages experienced by


vapor deposition of ethanol as carbon source
Source: CNTs were prepared using chemical

Note: TNTs are defined as titanate nanotubes produced through hydrothermal treatment.

the cationic doping materials. Titanate nanotubes doped with non-


metal elements show positive effect in catalytic performance.
Preparation methods, modifier source

Source: 2, 4 and 8 wt.% pentahydrated

Source: Lead(II) nitrate and Cd(NO3 )2

Doping titanate nanotubes with non-metal ions including N


Electrospinning, carbonization, and

Source: CNFs were prepared using


Hydrothermal, wet-impregnation

and exposed to H2 S atmosphere

[103], S [105] and C [18] have attracted considerable attentions


due to the increasing photocatalytic activity performed under
ammonium paratungstate

visible light irradiation. In 2001, Asahi et al. [116] found that


nitrogen-doped TiO2 nanomaterials exhibited superior photocat-
alytic activity among other non-metals (i.e. C, N, P, F and S) doped
Co-hydrothermal

Co-hydrothermal

TiO2 materials in various reactions. N element is the most inves-


hydrothermal

tigated dopant for narrowing the band gap energy of titanate


graphene

nanotubes as photocatalysts. N-doped TiO2 could be obtained by


adding different nitrogen sources such as ammonia, ammonium
salts or urea [103,116]. The reason for the improvement of cat-
alytic activity has been often related to the decrease of the band
gap energy of TiO2 , which is due to either N 2p states mixed with
Hybrid with carbon nanomaterials

O 2p states or the creation of an isolated impurity level above


Lead(II) sulfide (PbS) and CdS

valence band [79,116]. In general, substitutional type of N-doping


Carbon nano fibers (CNFs)

accompanied by oxygen vacancy is favored when the doping is car-


Carbon nanotubes (CNTs)
Tungsten trioxide (WO3 )

ried out under highly reducing condition; whereas interstitial type


N-doping takes place under oxygen-rich condition [103]. Substi-
tutional (N Ti O and Ti O N) and interstitial (characteristic NO)
Table 6 (Continued)

states created localized states above the valence band, resulting in


Graphene

visible light response and faster photocatalytic degradation.


Modifier

It should be bear in mind that the content of non-metal could


decease during annealing process, thereby, reducing its catalytic
activity. Hence, co-doping materials have been investigated to
Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142 141

compensate the drawbacks of each single doped titanate nano- nanotubes, factors affecting the formation of titanate nanotubes by
tubes. For example, co-doping titanate nanotubes such as Gd3+ N hydrothermal treatment, crystal structure, mechanism, structural-
[104], Pt N [102], and B N [117] are proposed to be more advan- activity correlation and modification of TiO2 -based nanotubes.
tageous over single nitrogen doping. Pt is used as electron trapper Template-assisted method allows the fabrication of materials with
[102], Gd3+ acted as a sensitizer to increase light absorption [104], regular and controlled morphology by adjusting the morphology
while B could minimize the oxygen vacancies produced by N doping of the applied template. However, it encounters difficulties such
[117]. Consequently, photo-charge separation, electron transfer as time consuming due to pre-fabrication and post-removal tem-
ability and photocatalytic activity can be enhanced as well as main- plate and contamination may occur during post-removal template.
taining non-metal doping stability in the surface of photocatalysts. Electrochemical anodization method can build highly ordered self-
organized TiO2 nanotubes immobilized on a titanium foil surface
6.3. Coupled semiconductors/binary composites with controllable pore size. However, this preparation method can
be considered as an expensive technique for large scale produc-
It has been reported that good matching of the conduction and tion of TiO2 nanotubes and length distribution of tubes is not
valence bands of two semiconductors could enable efficient charge well-defined over a large surface area. Meanwhile, TiO2 -based
carrier transfer from one to another [106]. Wide band gap energy nanotube prepared by alkaline hydrothermal method is a sim-
of titanate nanotubes coupling with small band gap semiconductor ple and environmental friendly synthesis technique as compared
could result in the formation of heterojunctions [107]. This allows to template-assisted and electrochemical anodization. Additional
electron to be injected from small band gap of the semiconductor advantages are more cost-effectiveness and larger amount of
(sensitizer) to its neighbor counterpart (titanate nanotubes) and TiO2 -based nanotubes can be produced in one-pot hydrothermal
improve separation of electron–hole pairs. process.
The modification of titanate nanotubes with quantum dots of TiO2 -based catalyst has become popular to be used as photo-
narrow band gap semiconducting materials such as CdS [108] and catalyst in the water treatment industry owning to its low cost,
PbS [108] is more promising as couple nanomaterials since it can high safety and high photocatalytic activity. However, the lack of
improve the photocatalytic activity effectively. CdS with a narrow visible light activity limits its practical applications. Hence, the
band gap of 2.4 eV allows CdS to act as photo-sensitizer to absorb improvement in the photocatalytic activity of TiO2 -based nano-
visible light irradiation [118]. The higher conduction band of CdS of tubes has been achieved via deposition with metal nanoparticles,
about 0.5 eV with respect to that of TiO2 could also enhance charge single- or co-doping with metal ions/non-metal ions, coupled with
separation [118]. Owning to the photo-generated electron transfer other semiconductors and hybridized with carbon nanomaterials.
from CdS to TiO2 , this charge separation restrains charge recom- The modification on TiO2 nanotubes exerts a substantial influence
bination. Thus, the coupling of CdS/titanate nanotubes has been in modifying the optical, interface and bulk properties of the pure
applied to enhance the photocatalytic degradation organic pol- TiO2 nanotubes, which then reflects to the charge carrier separa-
lutants [108]. In addition, the incorporation of WO3 into titanate tion and transfer behavior. In short, modified TiO2 -based nanotubes
nanotubes has also gained much attention due to smaller band have higher potential to be utilized as photocatalyst in large scale
gap energy of WO3 (2.8 eV) and higher corrosion stability in aque- applications.
ous solution [107]. Both the valence band and conduction band of
WO3 are lower than those of titanate nanotubes. Photo-generated
Acknowledgement
electrons would transfer from titanate nanotubes to that of WO3
and the holes transferred from WO3 to that of titanate nanotubes.
The authors would like to acknowledge The University of
Besides, titanate nanotubes composited with wide band gap energy
Malaya, Kuala Lumpur, Malaysia for the financial support under
of semiconductor such as ZnO (3.2 eV) [106] could suppress the
UM.C/HIR/MOHE/ENG/06 (D000006-16001).
charge recombination for higher photocatalytic activity.

6.4. Hybrid with carbon nanomaterials References

[1] M. Plodinec, A. Gajović, G. Jakša, K. Žagar, M. Čeh, J. Alloys Compd. 591 (2014)
Recently, titanate nanotubes composites with carbon nano-
147–155.
tubes, graphene and fullerene have attracted increasing attention. [2] Y. Tang, Z. Jiang, Q. Tay, J. Deng, Y. Lai, D. Gong, Z. Dong, Z. Chen, RSC Adv. 2
Carbon nanotubes show excellent mechanical strength, thermal (2012) 9406–9414.
conductivity, unique electronic properties and thermal stability [3] N. Pugazhenthiran, S. Murugesan, S. Anandan, J. Hazard. Mater. 263 (Pt 2)
(2013) 541–549.
[110]. Meanwhile, graphene with zero band gap possesses excel- [4] A. Ghicov, P. Schmuki, Chem. Commun. (2009) 2791–2808.
lent electrical and optical properties, large specific surface area and [5] L.L. Costa, A.G.S. Prado, J. Photochem. Photobiol. A 201 (2009) 45–49.
outstanding mechanical properties [111]. In addition, graphene can [6] H. Yu, J. Yu, B. Cheng, J. Lin, J. Hazard. Mater. 147 (2007) 581–587.
[7] P. Roy, D. Kim, K. Lee, E. Spiecker, P. Schmuki, Nanoscale 2 (2010) 45–59.
act as photo-sensitizer to transfer electron to the titanate nano- [8] A. El Ruby Mohamed, S. Rohani, Energy Environ. Sci. 4 (2011) 1065–1086.
tubes surface due to the more positive Fermi level of graphene [9] P. Roy, S. Berger, P. Schmuki, Angew. Chem. Int. Ed. 50 (2011) 2904–2939.
[119]. Solid fullerenes in the form of C60 with band gap energy of [10] D.V. Bavykin, J.M. Friedrich, F.C. Walsh, Adv. Mater. 18 (2006) 2807–2824.
[11] K.-C. Huang, S.-H. Chien, Appl. Catal. B 140/141 (2013) 283–288.
1.6–1.9 eV and a unique structure of delocalized ␲ electrons may act [12] T. Sekino, in: T. Kijima (Ed.), Inorganic and Metallic Nanotubular Materi-
as a sensitizer when attached to the surface of titanate nanotubes als: Recent Technologies and Applications, Springer-Verlag, Berlin, Germany,
[120]. These outstanding structural characters enable them to be 2010, pp. 17–32.
[13] J. Chen, H. Wang, X. Wei, L. Zhu, Mater. Res. Bull. 47 (2012) 3747–3752.
used as promising materials for environmental catalysts. In short,
[14] R.P. Antony, T. Mathews, C. Ramesh, N. Murugesan, A. Dasgupta, S. Dhara, S.
these hybrid materials showed large surface area with high quality Dash, A.K. Tyagi, Int. J. Hydrogen Energy 37 (2012) 8268–8276.
of active sites and able to inhibit the recombination electron–hole [15] R.-a. Doong, S.-m. Chang, C.-w. Tsai, Appl. Catal. B 129 (2013) 48–55.
[16] J. Yan, F. Zhou, J. Mater. Chem. 21 (2011) 9406–9418.
pairs.
[17] Y. Ohsaki, N. Masaki, T. Kitamura, Y. Wada, T. Okamoto, T. Sekino, K. Niihara,
S. Yanagida, Phys. Chem. Chem. Phys. 7 (2005) 4157–4163.
7. Conclusions [18] E.M. Neville, J.M.D. MacElroy, K.R. Thampi, J.A. Sullivan, J. Photochem. Photo-
biol. A 267 (2013) 17–24.
[19] H.-H. Ou, S.-L. Lo, Sep. Purif. Technol. 58 (2007) 179–191.
A series of comprehensive review on TiO2 -based nanotubes are [20] C. Bae, H. Yoo, S. Kim, K. Lee, J. Kim, M.M. Sung, H. Shin, Chem. Mater. 20 (2008)
discussed, which include three main routes used to synthesize TiO2 756–767.
142 Y.L. Pang et al. / Applied Catalysis A: General 481 (2014) 127–142

[21] J. Qiu, X. Li, X. Gao, X. Gan, B. Weng, L. Li, Z. Yuan, Z. Shi, Y.H. Hwang, J. Mater. [66] H.-K. Seo, G.-S. Kim, S.G. Ansari, Y.-S. Kim, H.-S. Shin, K.-H. Shim, E.-K. Suh,
Chem. 22 (2012) 23411–23417. Sol. Energy Mater. Sol. Cells 92 (2008) 1533–1539.
[22] J.H. Kim, X.H. Zhang, J.D. Kim, H.M. Park, S.B. Lee, J.W. Yi, S.I. Jung, J. Solid State [67] Y. Lan, X. Gao, H. Zhu, Z. Zheng, T. Yan, F. Wu, S.P. Ringer, D. Song, Adv. Funct.
Chem. 196 (2012) 435–440. Mater. 15 (2005) 1310–1318.
[23] A.M. Md Jani, D. Losic, N.H. Voelcker, Prog. Mater. Sci. 58 (2013) 636–704. [68] E.K. Ylhäinen, M.R. Nunes, A.J. Silvestre, O.C. Monteiro, J. Mater. Sci. 47 (2012)
[24] P. Hoyer, Langmuir 12 (1996) 1411–1413. 4305–4312.
[25] J.H. Lee, I.C. Leu, M.C. Hsu, Y.W. Chung, M.H. Hon, J. Phys. Chem. B 109 (2005) [69] G. Dawson, W. Chen, T. Zhang, Z. Chen, X. Cheng, Solid State Sci. 12 (2010)
13056–13059. 2170.
[26] C. Bae, Y. Yoon, H. Yoo, D. Han, J. Cho, B.H. Lee, M.M. Sung, M. Lee, J. Kim, H. [70] A.L. Papa, N. Millot, L. Saviot, R. Chassagnon, O. Heintz, J. Phys. Chem. C 113
Shin, Chem. Mater. 21 (2009) 2574–2576. (2009) 12682–12689.
[27] L. Yuan, S. Meng, Y. Zhou, Z. Yue, J. Mater. Chem. A 1 (2013) 2552–2557. [71] L. Miao, S. Tanemura, T. Jiang, M. Tanemura, K. Yoshida, N. Tanaka, G. Xu,
[28] M. Karaman, F. Sarıipek, Ö. Köysüren, H.B. Yıldız, Appl. Surf. Sci. 283 (2013) Superlattices Microstruct. 46 (2009) 357–364.
993–998. [72] D.V. Bavykin, V.N. Parmon, A.A. Lapkin, F.C. Walsh, J. Mater. Chem. 14 (2004)
[29] M. Müllner, T. Lunkenbein, M. Schieder, A.H. Gröschel, N. Miyajima, 3370–3377.
M. Förtsch, J. Breu, F. Caruso, A.H.E. Müller, Macromolecules 45 (2012) [73] D.V. Bavykin, B.A. Cressey, F.C. Walsh, Aust. J. Chem. 60 (2007) 95–98.
6981–6988. [74] D.V. Bavykin, A.N. Kulak, F.C. Walsh, Cryst. Growth Des. 10 (2010) 4421–4427.
[30] S.J. Ku, G.C. Jo, C.H. Bak, S.M. Kim, Y.R. Shin, K.H. Kim, S.H. Kwon, J.B. Kim, [75] G. Hasegawa, K. Kanamori, Y. Sugawara, Y. Ikuhara, K. Nakanishi, J. Colloid
Nanotechnology 24 (2013) 085308. Interface Sci. 374 (2012) 291–296.
[31] X. Hou, F. Li, G. He, J. Zhang, Ceram. Int. 40 (2014) 5811–5815. [76] S. Sreekantan, L.C. Wei, J. Alloys Compd. 490 (2010) 436–442.
[32] X. Meng, M.N. Banis, D. Geng, X. Li, Y. Zhang, R. Li, H. Abou-Rachid, X. Sun, [77] R. Ma, K. Fukuda, T. Sasaki, M. Osada, Y. Bando, J. Phys. Chem. B 109 (2005)
Appl. Surf. Sci. 266 (2013) 132–140. 6210–6214.
[33] D. Gong, C.A. Grimes, O.K. Varghese, W. Hu, R.S. Singh, Z. Chen, E.C. Dickey, J. [78] Y. Ma, Y. Lin, X. Xiao, X. Zhou, X. Li, Mater. Res. Bull. 41 (2006) 237–243.
Mater. Res. 16 (2001) 3331–3334. [79] X. Chen, S.S. Mao, Chem. Rev. 107 (2007) 2891–2959.
[34] S. Sreekantan, Z. Lockman, R. Hazan, M. Tasbihi, L.K. Tong, A.R. Mohamed, J. [80] C.T. Nam, J.L. Falconer, L.M. Duc, W.-D. Yang, Mater. Res. Bull. 51 (2014) 49–55.
Alloys Compd. 485 (2009) 478–483. [81] K. Das, S.K. Panda, S. Chaudhuri, J. Cryst. Growth 310 (2008) 3792–3799.
[35] E. Şennik, Z. Çolak, N. Kılınç, Z.Z. Öztürk, Int. J. Hydrogen Energy 35 (2010) [82] B. Santara, P.K. Giri, Mater. Chem. Phys. 137 (2013) 928–936.
4420–4427. [83] M.A. Khan, H.T. Jung, O.B. Yang, J. Phys. Chem. B 110 (2006) 6626–6630.
[36] N. Baram, Y. Ein-Eli, J. Phys. Chem. C 114 (2010) 9781–9790. [84] T. Kasuga, M. Hiramatsu, A. Hoson, T. Sekino, K. Niihara, Adv. Mater. 11 (1999)
[37] I. Roman, R.D. Trusca, M.-L. Soare, C. Fratila, E. Krasicka-Cydzik, M.-S. Stan, A. 1307–1311.
Dinischiotu, Mater. Sci. Eng. C 37 (2014) 374–382. [85] M. Zhang, Z. Jin, J. Zhang, X. Guo, J. Yang, W. Li, X. Wang, Z. Zhang, J. Mol. Catal.
[38] M. Nischk, P. Mazierski, M. Gazda, A. Zaleska, Appl. Catal. B 144 (2014) A: Chem. 217 (2004) 203–210.
674–685. [86] X. Pan, M.Q. Yang, X. Fu, N. Zhang, Y.J. Xu, Nanoscale 5 (2013) 3601–3614.
[39] H. Jha, P. Roy, R. Hahn, I. Paramasivam, P. Schmuki, Electrochem. Commun. [87] R. Ma, Y. Bando, T. Sasaki, Chem. Phys. Lett. 380 (2003) 577–582.
13 (2011) 302–305. [88] J.N. Nian, H. Teng, J. Phys. Chem. B 110 (2006) 4193–4198.
[40] V. Zwilling, E. Darque-Ceretti, A. Boutry-Forveille, D. David, M.Y. Perrin, M. [89] C. Huang, K. Zhu, M. Qi, Y. Zhuang, C. Cheng, J. Phys. Chem. Solids 73 (2012)
Aucouturier, Surf. Interface Anal. 27 (1999) 629–637. 757–761.
[41] M. Paulose, K. Shankar, S. Yoriya, H.E. Prakasam, O.K. Varghese, G.K. Mor, [90] P.-T. Hsiao, K.-P. Wang, C.-W. Cheng, H. Teng, J. Photochem. Photobiol. A 188
T.A. Latempa, A. Fitzgerald, C.A. Grimes, J. Phys. Chem. B 110 (2006) 16179– (2007) 19–24.
16184. [91] Y. Mao, S.S. Wong, J. Am. Chem. Soc. 128 (2006) 8217–8226.
[42] T. Ruff, R. Hahn, P. Schmuki, Appl. Surf. Sci. 257 (2011) 8177–8181. [92] Y.Q. Wang, G.Q. Hu, X.F. Duan, H.L. Sun, Q.K. Xue, Chem. Phys. Lett. 365 (2002)
[43] J.M. Macak, S.P. Albu, P. Schmuki, Phys. Status Solidi RRL 1 (2007) 181–183. 427–431.
[44] D. Kim, A. Ghicov, S.P. Albu, P. Schmuki, J. Am. Chem. Soc. 130 (2008) [93] S. Zhang, L.M. Peng, Q. Chen, G.H. Du, G. Dawson, W.Z. Zhou, Phys. Rev. Lett.
16454–16455. 91 (2003) 2561031–2561034.
[45] S.P. Albu, D. Kim, P. Schmuki, Angew. Chem. Int. Ed. 47 (2008) 1916–1919. [94] Á. Kukovecz, M. Hodos, E. Horváth, G. Radnóczi, Z. Kónya, I. Kiricsi, J. Phys.
[46] Y. Ji, K.-C. Lin, H. Zheng, J.-j. Zhu, A.C.S. Samia, Electrochem. Commun. 13 Chem. B 109 (2005) 17781–17783.
(2011) 1013–1015. [95] E. Morgado Jr., M.A.S. De Abreu, G.T. Moure, B.A. Marinkovic, P.M. Jardim, A.S.
[47] Q.A. Nguyen, Y.V. Bhargava, T.M. Devine, Electrochem. Commun. 10 (2008) Araujo, Chem. Mater. 19 (2007) 665–676.
471–475. [96] H.Y. Zhu, Y. Lan, X.P. Gao, S.P. Ringer, Z.F. Zheng, D.Y. Song, J.C. Zhao, J. Am.
[48] M.Y.A. Mollah, P. Morkovsky, J.A.G. Gomes, M. Kesmez, J. Parga, D.L. Cocke, J. Chem. Soc. 127 (2005) 6730–6736.
Hazard. Mater. 114 (2004) 199–210. [97] M.J. Li, Z.Y. Chi, Y.C. Wu, J. Am. Chem. Soc. 95 (2012) 3297–3304.
[49] E. Horváth, Á. Kukovecz, Z. Kónya, I. Kiricsi, Chem. Mater. 19 (2007) 927–931. [98] C.-K. Lee, C.-C. Wang, M.-D. Lyu, L.-C. Juang, S.-S. Liu, S.-H. Hung, J. Colloid
[50] Z.-Y. Yuan, B.-L. Su, Colloids Surf. A 241 (2004) 173–183. Interface Sci. 316 (2007) 562–569.
[51] T. Kasuga, M. Hiramatsu, A. Hoson, T. Sekino, K. Niihara, Langmuir 14 (1998) [99] M. Qamar, C.R. Yoon, H.J. Oh, N.H. Lee, K. Park, D.H. Kim, K.S. Lee, W.J. Lee, S.J.
3160–3163. Kim, Catal. Today 131 (2008) 3–14.
[52] H. Kochkar, N. Lakhdhar, G. Berhault, M. Bausach, G. Abdelhamid, J. Phys. [100] Q. Zhao, M. Li, J. Chu, T. Jiang, H. Yin, Appl. Surf. Sci. 255 (2009) 3773–3778.
Chem. C 113 (2009) 1672–1679. [101] R. Yuan, B. Zhou, D. Hua, C. Shi, J. Hazard. Mater. 262 (2013) 527–538.
[53] E. Morgado Jr., M.A.S. de Abreu, O.R.C. Pravia, B.A. Marinkovic, P.M. Jardim, [102] D. Slamet, Tristantini, M. Valentina, Ibadurrohman, Int. J. Energy Res. 37
F.C. Rizzo, A.S. Araújo, Solid State Sci. 8 (2006) 888–900. (2013) 1372–1381.
[54] S. Preda, V.S. Teodorescu, A.M. Musuc, C. Andronescu, M. Zaharescu, J. Mater. [103] C.-C. Hu, T.-C. Hsu, S.-Y. Lu, Appl. Surf. Sci. 280 (2013) 171–178.
Res. 28 (2013) 294–303. [104] H. Liu, G. Liu, G. Xie, M. Zhang, Z. Hou, Z. He, Appl. Surf. Sci. 257 (2011)
[55] D.L. Morgan, G. Triani, M.G. Blackford, N.A. Raftery, R.L. Frost, E.R. Waclawik, 3728–3732.
J. Mater. Sci. 46 (2011) 548–557. [105] F. Dong, W. Zhao, Z. Wu, Nanotechnology 19 (2008) 365607.
[56] L.M. Sikhwivhilu, S. Sinha Ray, N.J. Coville, Appl. Phys. A: Mater. Sci. Process. [106] L.S. Wang, M.W. Xiao, X.J. Huang, Y.D. Wu, J. Hazard. Mater. 161 (2009) 49–54.
94 (2009) 963–973. [107] M. Grandcolas, T. Cottineau, A. Louvet, N. Keller, V. Keller, Appl. Catal. B
[57] N. Viriya-empikul, T. Charinpanitkul, N. Sano, A. Soottitantawat, T. Kikuchi, 138/139 (2013) 128–140.
K. Faungnawakij, W. Tanthapanichakoon, Mater. Chem. Phys. 118 (2009) [108] Q. Li, T. Kako, J. Ye, J. Mater. Chem. 20 (2010) 10187–10192.
254–258. [109] S. Kim, M. Kim, Y.K. Kim, S.-H. Hwang, S.K. Lim, Appl. Catal. B 148/149 (2014)
[58] A. Turki, H. Kochkar, C. Guillard, G. Berhault, A. Ghorbel, Appl. Catal. B 138/139 170–176.
(2013) 401–415. [110] T. Jiang, L. Zhang, M. Ji, Q. Wang, Q. Zhao, X. Fu, H. Yin, Particuology 11 (2013)
[59] M. Kitano, E. Wada, K. Nakajima, S. Hayashi, S. Miyazaki, H. Kobayashi, M. 737–742.
Hara, Chem. Mater. 25 (2013) 385–393. [111] Z. Qianqian, B. Tang, H. Guoxin, J. Hazard. Mater. 198 (2011) 78–86.
[60] Y. Zhu, H. Li, Y. Koltypin, Y.R. Hacohen, A. Gedanken, Chem. Commun. (2001) [112] R. Marschall, L. Wang, Catal. Today 225 (2014) 111–135.
2616–2617. [113] Y. Yacin, M. Kilic, Z. Cinar, Appl. Catal. B 99 (2010) 469–477.
[61] H.H. Ou, C.H. Liao, Y.H. Liou, J.H. Hong, S.L. Lo, Environ. Sci. Technol. 42 (2008) [114] J. Yu, Q. Xiang, M. Zhou, Appl. Catal. B 90 (2009) 595–602.
4507–4512. [115] Y.L. Pang, A.Z. Abdullah, J. Hazard. Mater. 235/236 (2012) 326–335.
[62] B. Liu, E.S. Aydil, J. Am. Chem. Soc. 131 (2009) 3985–3990. [116] R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Science 293 (2001) 269–271.
[63] E. Morgado Jr., M.A.S. de Abreu, G.T. Moure, B.A. Marinkovic, P.M. Jardim, A.S. [117] P. Dong, Y. Wang, B. Liu, L. Guo, Y. Huang, S. Yin, J. Am. Ceram. Soc. 95 (2012)
Araujo, Mater. Res. Bull. 42 (2007) 1748–1760. 82–84.
[64] J. Huang, Y. Cao, Z. Deng, H. Tong, J. Solid State Chem. 184 (2011) 712–719. [118] R. Pan, Y. Wu, Z. Li, Z. Fang, Appl. Surf. Sci. 292 (2014) 886–891.
[65] R. Camposeco, S. Castillo, I. Mejia-Centeno, J. Navarrete, R. Gómez, Mater. [119] G. Hu, B. Tang, Mater. Chem. Phys. 138 (2013) 608–614.
Charact. 90 (2014) 113–120. [120] M. Grandcolas, J. Ye, K. Miyazawa, Ceram. Int. 40 (2014) 1297–1302.

You might also like