You are on page 1of 69

Accepted Manuscript

Hydrothermal/Solvothermal synthesis and treatment of TiO2 for photocatalytic


degradation of air pollutants: Preparation, characterization, properties, and
performance

Alireza Haghighat Mamaghani, Fariborz Haghighat, Chang-Seo Lee

PII: S0045-6535(18)32351-8

DOI: 10.1016/j.chemosphere.2018.12.029

Reference: CHEM 22728

To appear in: Chemosphere

Received Date: 16 July 2018

Accepted Date: 05 December 2018

Please cite this article as: Alireza Haghighat Mamaghani, Fariborz Haghighat, Chang-Seo Lee,
Hydrothermal/Solvothermal synthesis and treatment of TiO2 for photocatalytic degradation of air
pollutants: Preparation, characterization, properties, and performance, Chemosphere (2018), doi:
10.1016/j.chemosphere.2018.12.029

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form.
Please note that during the production process errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Hydrothermal/Solvothermal synthesis and treatment of TiO2 for photocatalytic


degradation of air pollutants: Preparation, characterization, properties, and
performance

Alireza Haghighat Mamaghani, Fariborz Haghighat*, Chang-Seo Lee

Department of Building, Civil and Environmental Engineering, Concordia University,


Montreal, Canada

Abstract

Photocatalytic oxidation (PCO) is a well-known technology for air purification and has been

extensively studied for removal of many air pollutants. Titanium dioxide (TiO2) is the most

investigated photocatalyst in the field of environmental remediation owed to its chemical

stability, non-toxicity, and suitable positions of valence and conduction bands. Various

preparation techniques including sol-gel, flame hydrolysis, water-in-oil microemulsion,

chemical vapor deposition, solvothermal, and hydrothermal have been employed to obtain

TiO2 materials. Hydro-/Solvothermal (HST) synthesis, focus of the present work, can be

defined as a preparation method in which crystal growth occurs in a solvent at relatively low

temperature (< 200 ˚C) and above atmospheric pressure. This paper aims to provide a

comprehensive and critical review of current knowledge regarding the application of HST

synthesis for fabrication of TiO2 nanostructures for indoor air purification. TiO2

nanostructures are categorized from the morphological standpoint (e.g. nanoparticles,

nanotubes, nanosheets, and hierarchically porous) and discussed in detail. The influence of

preparation parameters including hydrothermal time, temperature, pH of the reaction

medium, solvent, and calcination temperature on physical, chemical, and optical properties of

TiO2 is reviewed. Considering the complex interplay among catalyst properties, a special

*Corresponding author, fariborz.haghighat@concordia.ca


E-mail addresses: alireza.haghighatmamaghani@mail.concordia.ca (A. Haghighat M), chang-
seo.lee@concordia.ca (C.S. Lee)

1
ACCEPTED MANUSCRIPT

emphasis is placed on elucidating the interconnection between various photocatalyst features

and their impacts on photocatalytic activity.

Keywords: Titanium dioxide; Photocatalytic oxidation; Hydrothermal; Solvothermal;


Morphology; Characterization.

List of acronyms
BET Brunauer-Emmett-Teller
BTEX Benzene, toluene, ethyl benzene, o-xylene
CB conduction band
DSC differential scanning calorimetry
DRIFTS diffuse reflectance infrared Fourier transform spectroscopy
e--h+ electron-hole pair
EDX energy dispersive X-ray
EPR electron paramagnetic resonance
ESR electron spin resonance
FTIR Fourier transform infrared
HST Hydro-/solvothermal
N2 ads-des nitrogen adsorption-desorption
OH surface hydroxyl group
OH• hydroxyl radical
O2•- superoxide radical anion
P25 Aeroxide (Evonik), mixed anatase and rutile TiO2
PCO photocatalytic oxidation
PL photoluminescence spectroscopy
SEM scanning electron microscopy
SPR surface plasmon resonance
TBOT titanium butoxide
TCE trichloroethylene
TEM transmission electron microscopy
TFA trifluoroacetic acid
TGA thermal gravimetric analysis
TiNTs titanate nanotubes
TIP titanium isopropoxide
TiO2 titanium dioxide
TNTs TiO2 nanotubes
VB valence band
VOC volatile organic compound
XPS X-ray photoelectron spectroscopy
XRD X-ray diffraction

2
ACCEPTED MANUSCRIPT

1. Introduction

Current levels of indoor air pollutants are expected to escalate in future considering the

decline in outdoor air quality and tighter building envelopes. Consequently, concerns

regarding the impact of indoor air quality on human health especially in well-insulated

buildings have created a pressing need for reliable air cleaning technologies (Verbruggen,

2015). In this regard, photocatalytic oxidation has been regarded as an effective and

economically viable option to remove air pollutants at low concentrations (Demeestere et al.,

2007; Sánchez et al., 2012; Zhong and Haghighat, 2015; Mamaghani et al., 2018b).

PCO has the capability of degrading air pollutants (such as volatile organic compounds

(VOCs)) with low concentrations at room temperature and produces innocuous CO2 and H2O

as the main final products. Excitation of electrons from the valence band (VB) to the

conduction band (CB) by photons having energy greater than semiconductor’s band gap is

the underlying process in PCO (Nakata and Fujishima, 2012). This phenomenon leads to the

creation of photogenerated charge carriers (electrons in the CB and holes in the VB) which

then react with adsorbed water, surface hydroxyl (OH) groups, and oxygen and form highly

reactive oxidizing species such as hydroxyl radical (OH•) and superoxide radical anion (O2•-).
+
(
𝑇𝑖𝑂2 + ℎ𝑣→ 𝑇𝑖𝑂2 𝑒𝐶𝐵 + ℎ

𝑉𝐵
) (1)

( +)
𝑇𝑖𝑂2 ℎ𝑉𝐵 + 𝐻2𝑂→𝑇𝑖𝑂2 + 𝐻
+
+ 𝑂𝐻

(2)

( +)
𝑇𝑖𝑂2 ℎ𝑉𝐵 + 𝑂𝐻 →𝑇𝑖𝑂2 + 𝑂𝐻

(3)

( ‒)
𝑇𝑖𝑂2 𝑒𝐶𝐵 + 𝑂2→𝑇𝑖𝑂2 + 𝑂
•‒
2 (4)

Inevitably, a part of e--h+ pairs recombine in the bulk or on the surface without participating

in photochemical reactions and simply dissipate heat. Similar to heterogeneous catalysis,

transfer of pollutant molecules from air onto the surface of titanium dioxide is essential in

photocatalysis. The mass transfer of pollutants entails the external and internal diffusions and

the adsorption on active sites. Chemical reactions between pollutant molecules adsorbed on

3
ACCEPTED MANUSCRIPT

the surface and active radicals break down large molecules to smaller ones and eventually to

CO2, H2O and light by-products (Xiao et al., 2015; Wan et al., 2018).

The performance of PCO processes (i.e. adsorption, electron excitation, reactions, etc)

strongly depends on operating parameters including relative humidity, airflow rate, challenge

compound type and concentration, and light source and intensity (Shayegan et al., 2018).

Among numerous semiconductor photocatalysts developed for PCO hitherto, TiO2 based

materials are presently the most promising photocatalysts for indoor air purification owed to

their unique features such as suitable band gap energy, desirable positions of VB and CB for

oxidation and reductions reactions, and activity towards many pollutants (Fujishima et al.,

2000; Fujishima et al., 2008). Theoretical and experimental studies have pointed to the fact

that the photocatalytic response is greatly affected by the properties of TiO2 (Demeestere et

al., 2007; Romero-Vargas Castrillón and de Lasa, 2007; Einaga et al., 2015). In particular,

crystallinity, phase structures, surface area, surface OH content, pore size distribution, band

gap, crystalline size, and the exposed surface facets were proven to exert significant influence

on PCO processes (Ismail and Bahnemann, 2011; Mamaghani et al., 2017). These textural,

chemical, and electronic properties of TiO2 can be precisely tailored by manipulating the

preparation conditions to optimize the performance and lifetime of photocatalyst.

TiO2 can be synthesized by numerous techniques including: sol-gel, hydrothermal,

solvothermal, pyrolysis, vapour deposition, chemical precipitation, micelles, and direct

oxidation (Chen and Mao, 2007; Ismail and Bahnemann, 2011). Hydrothermal synthesis is

generally defined as crystallization above the room temperature and at high pressure (> 1

atm) via heterogeneous reactions in aqueous/non-aqueous media (Byrappa and Adschiri,

2007; Hayashi and Hakuta, 2010). Most researchers have applied the term solvothermal when

chemical reactions take place in non-aqueous solvents. Hydrothermal synthesis is usually

performed in sealed stainless steel autoclave with Teflon liner, which can withstand high

4
ACCEPTED MANUSCRIPT

temperature and pressure associated with this process. The central reason for employing such

conditions is that materials that are insoluble or difficult to dissolve go into solution as

complexes owing to cooperative effect of temperature, pressure, and solvent. Therefore, the

properties of solvent (in terms of dielectric constant, viscosity, density, etc.) and solution (e.g.

solubility, stability, yield, and dissolution-precipitation reactions) under hydrothermal

conditions can affect the structure of final TiO2 products (Byrappa and Adschiri, 2007). A

number of features make hydrothermal synthesis an ideal method for preparation of TiO2:

defect-free nano-crystals with high specific surface area, low agglomeration between

particles, good crystallinity, high product purity, crystal symmetry, narrow particle size

distribution, formation of anatase at relatively low temperature (< 200 ˚C), low energy

consumption, and inexpensive instrumentation (Yu et al., 2006a; Byrappa and Adschiri,

2007; Zhu et al., 2011; Grabowska et al., 2016). More importantly, the key characteristics of

TiO2 can be tuned to a great extent by systematically adjusting the experimental parameters

including: hydrothermal temperature, duration, pressure (percentage fill), solvent, pH of

solution, mineralizer, surfactant, type of acid/base, calcination temperature, and the type of

titanium precursor.

Realizing that many reviews are available on introduction to TiO2 and its preparation

methods (Ismail and Bahnemann, 2011; Bian et al., 2016; Mamaghani et al., 2017), in this

article a comprehensive and focused review of the attempts made to synthesize TiO2 via HST

method for indoor air treatment is provided. Here, the focus is mainly on the synthesis

procedure, characterization, properties, and photocatalytic activity of TiO2 based materials to

further the understanding of HST method capabilities and to highlight the research gaps need

to be addressed.

5
ACCEPTED MANUSCRIPT

2. Hydro-/solvothermally prepared TiO2 photocatalysts with different

morphologies

Structural dimensionality of TiO2 has a great impact on its properties and photocatalytic

activity towards air contaminants. TiO2 photocatalysts can be divided into four main

categories based on their dimensionality as illustrated in Fig. 1 (Nakata and Fujishima, 2012).

Numerous TiO2 morphologies, such as nanoparticles, microspheres, nanotubes, nanosheets,

hollow spheres, and hierarchically porous, have been fabricated and tested for environmental

decontamination (Nakata and Fujishima, 2012; Verbruggen, 2015). In the following sections,

each structural design is discussed in detail from the perspective of hydrothermal preparation

parameters, properties of TiO2, and the activity of TiO2 in PCO of indoor air pollutants.

Fig. 1. Morphological classification of TiO2 photocatalysts based on their


dimensionality; (a) (Wu et al., 2007), (b) (Hernández-Alonso et al., 2011), (c) (Wang et
al., 2015), and (d) (Yu et al., 2007d).

Table 1 presents significant works conducted on PCO of indoor air pollutants along with the

hydrothermal preparation parameters, research focus, key properties of the final product, and

improvement in efficiency compared to P25 (as a reference photocatalyst).

Table 1. Preparation conditions, key properties, and photocatalytic performance of


TiO2 based photocatalysts.

6
ACCEPTED MANUSCRIPT
Morphological Titanium Hydrothermal time and Post thermal treatment Crystal Surface Crystal Enhancement
Reaction media Research focus Characterization methods Pollutant Ref.
classification source temperature temperature and time phase area (m2/g) size (nm) factor a

Triton X-100 (surfactant); n- (van der Meulen


hexanol (co-surfactant); XRD, BET, TEM,
TBOT 13 h at 120 ˚C - Impact of the type of acid A, R, A+R 40-260 5-14 Propane 70 et al., 2007)
Cyclohexane; HCl/HNO3 (crystal DRIFTS, UV-vis
phase determining agent)

(Chen et al.,
NaOH; HNO3 (crystal phase Impact of HNO3
TBOT 24 h at 180 ˚C 50 ˚C XRD, N2 ads-des, PL A+R+B 92-137 7.5-23 Toluene 100 2015b)
determining agent) concentration

Drying at 100 ˚C for (Cano-Casanova


Ethanol-water (solvent); HCl 12 h followed by XRD, SEM, N2 ads-des, A+B,
TIP 12 h at 180 ˚C Impact of HCl concentration 100-135 8-24 Propene 43 et al., 2018)
(crystal phase determining agent) calcination at 350 ˚C FTIR, TGA, UV-vis A+B+R
for 2

110 ˚C overnight Impact of titanium precursor; (Deng et al.,


Water-Ethanol (solvent);
TBOT, followed by Impact of catalyst surface
Ammonia; HNO3 (phase 24 h at 180 ˚C XRD, BET, TEM, TGA A, R 19-165 8-65 Hexane 41 2002)
TiCl4 calcination at 250-550 sulfation; Impact of post-
composition controller); HCl
˚C for 3 h treatment calcination

110 ˚C overnight
Impact of HNO3
TBOT
Water-Ethanol (solvent)l; HNO3
24 h at 180 ˚C
followed by
concentration Impact of XRD, BET, TEM A, R, A+R 39-87 12-59
Hexane; 53 (hexane), (Wu et al., 2004)
(phase composition controller) calcination at 450 ˚C Methanol 12.5 ( methanol)
catalyst surface sulfation
for 3 h

NaOH (for synthesis of titanate (Chen et al.,


Nanoparticles/ Anatase nanotubes); HF, HCl, HNO3, Impact of type of acid and XRD, BET, SEM, FTIR, A, A+B,
24 h at 180 ˚C 50 ˚C 36-153 11-50 Toluene 113 2015a)
Microspheres powder H2SO4 and CH3COOH (phase concentration XPS A+R+B
(0-D) composition controller)

Ethanol; HNO3 (hydrolysis (Shi and Weng,


Impact of water to alcohol
TiCl3 catalyst); Urea ( hydrolysis 24 h at 80 ˚C 60 ˚C for 24 h XRD, BET, UV-vis A, R, A+R - 4.7-11.9 Formaldehyde 0 2008)
ratio in reaction solution
catalyst)

Impact of TIP and H2O (Maira et al.,


XRD, BET, TEM, UV-
TIP Water-Isopropanol (solvent) 8 h at 100-150 ˚C 65 ˚C for 24 h concentration, TIP addition A 20-270 2.3-27 TCE - 2000)
VIS, Raman spectra, TGA
rate

Comparison between
TBOT Water; Ethanol 12 h at 200 ˚C 450 ˚C for 3 h hydrothermal and thermal
TG, DSC, XRD, TEM,
A 115-148 7.2-9.8 Benzene -
(Wu et al., 2007)
HR-TEM, BET, SEM
treatments

Comparison between (Ardizzone et al.,


XRD, BET, UV-vis, XPS,
TIP 2-propanol; Water 500 h at 60 ˚C - hydrothermal and thermal A+B 145-200 6-8 Toluene 27 2008)
FTIR
treatments

Comparison between (Maira et al.,


Isopropanol (solvent in sol-gel
TIP 8 h at 200 ˚C 65 ˚C for 24 h hydrothermal and thermal XRD, BET, UV-vis, FTIR A 95 11 Toluene - 2001)
step); Ethanol
treatments

Comparison between (Ardizzone et al.,


XRD, BET, SEM, TEM,
TIP Water; Propanol (oil phase) 500 h at 60 ˚C - hydrothermal and thermal A+B 108-200 6-11 NOx - 2007)
XPS
treatments

7
ACCEPTED MANUSCRIPT (Zhou et al.,
80 ˚C for 12 h in Impact of hydrothermal XRD, SEM, TEM, N2 ads-
Ti(SO4)2 Water; CO(NH2)2 1-12 h at 160 ˚C A 107-196 12.7-22.5 Acetone 137 2010)
vacuum oven duration des

TiCl4 Na2SO4, CO(NH2)2, Ethanol 0.5-36 h at 140-200 ˚C 80 ˚C for 12 h


Influence of reaction time and SEM, TEM, XRD, N2 ads-
A
159.4–
2-7 Benzene 55
(Du et al., 2011)
temperature des 265.4

(Yu and Zhang,


Ethanol (control the hydrolysis); 80 ˚C for 10 h in Impact of the concentration XRD, SEM, TEM, N2 ads-
TBOT 12 h at 180 ˚C A, A+B 77-103 9.7-19.5 Acetone 130 2010)
KCl; NH4F (hollowing effect) vacuum oven of NH4F des

H2SO4 (pH adjuster); XRD, N2 ads-des, (Yu and Shi,


Impact of the concentration
Ti(SO4)2 Trifluoroacetic acid (hollowing 12 h at 180 ˚C 80 ˚C for 12 h SEM,TEM, XPS, FTIR, A 83-96 10.8-13.3 Acetone 80 2010)
of trifluoroacetic acid
effect) PL spectra
Vacuum drying (Liang et al.,
followed by Effect of calcination XRD, SEM, TEM, N2 ads-
Ti(SO4)2 NH4F (F source); H2O2 3 h at 180 ˚C A, A+R 3.2-20.8 78.5-97.7 Acetone - 2017)
calcination at 300- temperature des, UV-vis, XPS, EIS
1100 ˚C for 2 h

Graphite powder; H2SO4; KMnO4; 40 ˚C under vacuum XRD, TEM, XPS, Raman (Yu et al., 2018)
TBOT 12 h at 180 ℃ Effect of graphene A 62.5-105 - Formaldehyde -
H2O2; HCl; NaNO3 condition spectroscopy; UV-vis

(Roso et al.,
Graphene oxide; Ethanol; SEM, XRD, TGA, BET,
P25 3 h at 120 ˚C 70 ˚C for 12 h Effect of graphene A+R - - Methanol - 2015)
Polyacrylonirile scaffolds Raman spectroscopy

Acetic acid; Sodium dodecyl XRD, FTIR, TEM, SEM, (Yadav and Kim,
TIP sulfate; Ammonia; Graphene oxide; 4 h at 130 ˚C 70 ˚C Effect of graphene PL, EDS, Raman A - 10-15 Benzene - 2016)
Ethanol spectroscopy
(Ebrahimi and
Graphite powder; H3PO4; H2SO4; XRD, N2 ads-des, SEM,
P25 14 h at 150 ˚C - Effect of graphene A+R 70.2-73.2 30.8-36 Acetaldehyde 26 Fatemi, 2017)
HCl; H2O2; KMnO4; Ethanol TEM, FTIR, UV-vis

(Zhang et al.,
Ethanol-water (solvent); Graphite XRD, SEM, TEM, UV-
P25 24 h at 120 ˚C 60 ˚C Effect of graphene A+R - - Benzene 357 2010)
oxide vis, ESR

Calcined at 500 ˚C for SEM, TEM, XRD, TGA, (Suave et al.,


Hexadecylamine; Ethanol;
TIP 16 h at 130 ˚C 2 h under nitrogen Effect of graphene UV-vis, Raman A+B 120-130 7-10 NO -50 2017)
Potassium chloride; Graphite oxide
flow spectroscopy
(Wang et al.,
Water-Ethanol (solvent); XRD, Raman spectra, N2
TBOT 12 h at 180 ˚C 80 ˚C for 2 h Impact of graphene A 153.5-170 7.6-7.9 Acetone 160 2012)
Grapheme oxide ads-des, SEM, TEM, XPS

TiCl3
Hexamethylenetetramine (N
2 h at 190 ˚C 60 ˚C
Effect of rutile, brookite, and XRD, BET, UV-vis, XPS,
B, B+R 75.3-140.5 - NO 46
(Li et al., 2017)
source); g-C3N4; Ethanol (solvent) g-C3N4 FTIR, TEM, PL, EIS

XRD, N2 ads-des, UV-vis,


TiCl4
Ammonia; Ethanol; (NH4)2HPO4;
8 h at 190 ˚C 60 ˚C
Effect of hydroxyapatite
XPS, FTIR, TEM, TPD, A 54-98 11.2-39.9 NO -
(Yao et al., 2017)
Ca(NO3)2; support
EIS, ESR
Acetic acid; tert-butanol; HNO3; (Sboui et al.,
Raman spectra, SEM,
TBOT Commercial baking parchment 3 h at 120-140 ˚C 60 ˚C for 3 h Effect of support material A - - 2-propanol -34 2018)
UV-vis, XPS, TGA
paper
XRD, SEM, XPS, ICP- (Kibanova et al.,
Hectorite and kaolinite clays Effect of different clays as Toluene, D- 11 (toluene)
TIP 5 h at 180 ˚C 60 ˚C for 3 h OES, BET, porosimetry A 15.9-139.5 9.6 2009)
(support); HCl; Ethanol support material limonene -22 (limonene)
analysis
(Kibanova et al.,
Hectorite clay (support); HCl; Effect of Hectorite clay as
TIP 5 h at 180 ˚C 110 ˚C for 5 h N2 ads-des A - - Formaldehyde 70 2012)
Ethanol support material

8
ACCEPTED MANUSCRIPT (Jo and Kang,
Polyaniline (support); Ammonium Effect of Polyaniline as XRD, SEM, TGA, UV- 14 (B), 5 (T)
TiCl4 24 h at 95 ˚C 460 ˚C for 2 h A - 13.9-14.7 BTEX 2013)
sulfate; Urea; Ethanol support material vis, FTIR, EDX 7 (E), 4 (X)

HAuCl4 (Au source); Sodium 80 ˚C for 18 h in Impact of dopant SEM, TEM, XRD, N2 ads- (Yu et al., 2009b)
TBOT 7 h at 180 ˚C A 180-185 7-8.25 Formaldehyde 90
citrate vacuum oven concentration des, XPS, UV-vis, PL

KAuCl4, Pd(C5H7O2)2, H2PtCl6, Impact of various noble XRD, SEM, TEM, EDX, (Grabowska et al.,
TBOT AgNO3 (Au, Pd, Pt, Ag sources 4 h at 180 ˚C 40 ˚C metals anchored on the TiO2 N2 ads-des, DRS, XPS, A 235 - Toluene - 2016)
respectively) surface UV-vis, PL
XRD, BET, PL, FTIR, (Zeng et al.,
TBOT Ag nanowires; Ethanol; 10 h at 160 ˚C Vacuum drying Effect of Ag nanowires SEM, TEM, XPS, ESR, A 99.8-112.5 - Acetaldehyde - 2018)
EIS

TiF4
HAuCl4 (Au source); Sodium
48 h at 180 ˚C 80 ˚C
Impact of the core-shell
TEM, XRD, UV-vis A - - Acetaldehyde 18
(Wu et al., 2009)
citrate; Ascorbic acid structure

(Zhao et al.,
TiO2 ESR, photocurrent
NaOH; HNO3; NH4OH 48 h at 150 ˚C 450 ˚C for 2 h Impact of nitrogen doping - - - Toluene - 2012)
powder density, XPS

Drying at 100 °C
Urea (source of N); Ammonium Toluene
TBOT fluoride (sources of F); Ethanol 20 h at 150 ˚C
overnight and
Effect of N and F doping XRD, EDX, XPS, UV-vis A - 12.6-14.1 Ethyl benzene -
(Shin et al., 2015)
calcination at 400 °C
(solvent), HNO3 o-Xylene
for 3 h
Comparison between (Fresno et al.,
Ethanol; Ph3SnOH (Tin source); XRD, Raman spectra, Methyl
TiCl4 8 h at 150 ˚C - hydrothermal and thermal A, A+R, R 24-138 6-37 93 2005)
CH2Cl2; HCl BET, DRIFT cyclohexane
treatments; Impact of Sn
(Mattsson et al.,
Ammonia; Ethanol; Impact of dopant SEM, TEM, XRD, FTIR,
TIP 15 h at 200 ˚C - A 62 - Acetone - 2006)
Nb(OEt)5(niobium source) concentration UV-vis, XPS

(Shengwei et al.,
Ethanol; NH4HF2 (for surface Impact of surface XRD, TEM, N2 ads-des,
TBOT 10 h at 150 ˚C 80 ˚C for 2 h A+B, A 121-197 8.1-11.5 Acetone 250 2009)
fluorination and adjusting the pH) modification with NH4HF2 XPS, UV-vis, PL

120 ˚C overnight (Nguyen and Bai,


NaOH (for nanotube formation);
followed by Impact of post thermal 185 (NOx), 100
P25 HNO3 (post-treatment acid 24 h at 135 ˚C XRD, BET, SEM, TEM A 61-390 4.5-18 NOx; NO2 2014)
calcination at 400-600 treatment temperature (NO2)
washing)
˚C for 2 h

Impact of titanium precursor (Hernández-


NaOH (for nanotube formation); 100 ˚C overnight or Impact of type of acid
TiO2
HNO3 and HCl (post-treatment acid 24-72 h at 130-150 ˚C calcination at 250-500 Impact of hydrothermal time;
XRD, BET, TEM, EDX,
A, T 57-426 - TCE 10
Alonso et al.,
powder Raman spectra
washing) ˚C for 3 h Impact of post thermal 2011)
Nanotubes, treatment temperature
nanofibers,
nanowires
(1-D) 80 ˚C for 8 h followed
P25
KOH (for nanowire formation);
48 h at 150 ˚C by calcination at 300-
Impact of post thermal XRD, SEM, EDX, N2 ads- T, A, A+B,
5-200 4-311 Acetone 71
(Yu et al., 2007a)
HCl (post-treatment acid washing) treatment temperature des, UV-vis, TEM A+R, R
900 ˚C for 2 h

NaOH (for tubular shape


48 h at 150 ˚C followed 80 ˚C for 8 h in Impact of hydrothermal post XRD, SEM, TEM, N2 ads- (Yu et al., 2006c)
P25 formation), HCl (post-treatment A, T 90-356 6-30 Acetone 47
by 1-24 h at 200 ˚C vacuum oven treatment duration des
acid washing step)

9
ACCEPTED MANUSCRIPT
Impact of titanium precursor; (Chatterjee et al.,
NaOH (for tubular shape
TiO2 Impact of hydrothermal XRD, BET, FTIR, SEM,
formation), HCl (post-treatment 12 h at 100-200 ˚C - A, R, T 35-245 - Ethylene - 2010)
powder temperature; Impact of post TEM, EDX
acid washing step)
thermal treatment

NaOH (formation of tubular 110 ˚C overnight (Vijayan et al.,


Impact of dopant XRD, N2 ads des,
Anatase structure); HCl (post treatment acid followed by
48 h at 120 ˚C concentration on titania SEM,TEM, FTIR, UV- A 140-262 9.5-12 Acetaldehyde 330 2010)
powder washing step); Hexachloroplatinic calcination at 400 ˚C
nanotubes vis, EPR spectra
acid (Pt source) for 1 h

Drying at 90 ˚C for 12
Hexamethylene tetramine (N
TiCl3 source); NaOH; HCl (post 24 h at 150 ˚C
h followed by
Influence of nitrogen doping
XRD, SEM, TEM, UV-
A 82.4-93.2 - BTEX -
(Jo et al., 2014a)
calcination at 400 ˚C vis, N2 ads-des
preparation cleaning)
for 2 h

TBOT
Hexane (solvent); NaOH; Acetic
48 h at 160 ˚C 400 ˚C for 6 h
Impact of nanotube loading XRD, SEM, TEM, EDX,
A - - BTEX
65 (B), 61 (T), (Jo et al., 2014b)
acid on carbon fibers UV-vis 48 (E), 35 (X)

Carbon fiber; Hydrogen


Coupling titania nanowires XRD, TEM, SEM, N2 ads-
TBOT
tetrachloroaurate (III) hydrate
48 h at 110 ˚C 80 ˚C with carbon fiber and Au des, XPS, UV-vis, FTIR, A 46.6-78.5 - Styrene -
(Shi et al., 2017)
(HAuCl4·3H2O) (source of Au);
dopant PL
Ethanol

TBOT
HF (shape directing agent for
24 h at 190 ˚C 80 ˚C
Impact of surface fluorination XRD, TEM, SEM, N2 ads-
A 86-140 11.5-19.5 Ammonia 100
(Wu et al., 2014)
nanosheets) and morphology des, XPS, UV-vis, PL

Impact of the concentration (Xiang et al.,


HF (shape directing agent for SEM, TEM, XPS, N2 ads-
TBOT 24 h at 180 ˚C 80 ˚C for 6 h of HF in hydrothermal A 97-156 8.9-17.9 Acetone 700 2010)
nanosheets) des, PL
medium

Ethanol (solvent in nanoplates (Sofianou et al.,


Nanosheets fabrication); HF (capping agent); Impact of solvent type on XRD, XPS, SEM, TEM,
TIP 24 h at 180 ˚C 70 ˚C A 2-113 - NO - 2012)
(2-D) Water (solvent in microspheres architecture of final product BET
fabrication)

24 h at 180 ˚C for (Wang et al.,


HF (shape directing agent for
TBOT, nanosheets; 48 h at 200 Impact of TiO2 exposed XRD, BET, SEM, TEM,
nanosheets); KOH (shape directing 80 ˚C for 6 h A 38-52 - Toluene - 2015)
P25 ˚C followed by 3 h at facets and morphology Raman spectra, DRIFTS
agent for bipyramids)
200 ˚C for bipyramids

80 ˚C for 10 h in Impact of calcination


TBOT
HF (shape directing agent for
24 h at 200 ˚C
vacuum oven followed temperature during post XRD, SEM, TEM, XPS,
A, A+R 1.4-62 - Acetone 110
(Lv et al., 2011)
nanosheets) by calcination at 300- thermal treatment of N2 ads-des
1250 ˚C for 2 h. nanosheets

80 ˚C for 8 h in Impact of hydrothermal XRD, SEM, TEM, N2 ads- (Yu et al., 2007c)
TBOT Water 1-36 h at 180 ˚C A, A+B 150-460 4.2-9 Acetone 300
vacuum oven treatment duration des

10
ACCEPTED MANUSCRIPT
80 ˚C for 10 h in Impact of hydrothermal time XRD, N2 ads-des, SEM, (Yu et al., 2007b)
TBOT Water 1-24 h at 100-200 ˚C A 134-521 5.3-9.4 Acetone 267
vacuum oven and temperature TEM, XPS

HCl (pH adjuster in hydrothermal XRD, TEM, Raman


100 ˚C for 10 h in Impact of the pH of (Yu et al., 2006a)
TBOT solution); NH3.H2O (pH adjuster in 5 h at 180 ˚C spectroscopy, FTIR, N2 A 81-186 7.7-20.3 Acetone 128
vacuum oven hydrothermal solution
hydrothermal solution) ads-des, XPS, UV-vis

NaOH (creation of titanate 48 h at 150 ˚C for (Xiang and Yu,


nanotubes); HCl (post treatment synthesis of titanate SEM, TEM, XRD, N2 ads-
P25 80 ˚C for 6 h Impact of morphology A 9.7-39 - Acetone 10 2011)
acid washing step); Ethanol nanotubes; 24 h at 180 des, UV-vis
(solvent); HF (capping agent) ˚C

Porous and 55 ˚C overnight in


Decaoxyethylene cetyl ether (Xiang and Yu,
interconnected vacuum oven followed Impact of post thermal XRD, BET, SEM, TEM,
TIP surfactant (structural-directing 24 h at 80 ˚C A, R, A+R 2-201 5.8-41.3 Ethylene 68 2011)
architectures by calcination at 350- treatment temperature FTIR, UV-vis
agent)
(3-D) 800 ˚C for 4 h

(Cheng et al.,
Ethanol (solvent); CuSO4.5H2O
TBOT 24 h at 150 ˚C 40 ˚C for 4 h Effect of Cu2O XRD, SEM, TEM, UV-vis A - - Toluene 38 2018)
(Cu source); NaOH; Glucose

Impact of BiVO4 (Longo et al.,


Bi(NO3)3.5H2O (Bi source); HNO3;
TIP 24 h at 180 ˚C 100 ˚C for 2 h Comparison between thermal BET, XRD, XPS, UV-vis A 1.4-178 7-23 Isopropanol - 2014)
NaOH; NH4VO3 (V source)
and hydrothermal routes

Impact of different WO3 to (Puddu et al.,


Ammonium metatungstate (W
TiOSO4 12 h at 150 ˚C - TiO2 molar ratios in XRD, N2 ads-des, SEM A 7.6-136 18-30 TCE - 2007)
source); Ethanol
nanocomposite

Tetraethyl orthosilicate (Si source); 55 ˚C in vacuum oven Impact of second phase (SiO2 (Chen et al.,
Zirconium propoxide (Zr source); followed by or ZrO2) in TiO2 structure XRD, N2 ads-des, SEM,
TIP 24 h at 80 ˚C A 45-298 5.3-17.1 Ethylene - 2009)
Decaoxyethylene cetyl ether calcination at 350-800 Impact of calcination TEM, FTIR, UV-vis
(surfactant) ˚C for 4 h temperature

XRD, Raman spectra, N2


TIP
Pluronic P123; tetraethyl
24 h at 100 ˚C Calcination at 600 ˚C Effect of Ti/Si ratio ads-des, SEM, TEM, A+R 419-787 14.3-50.5 Formaldehyde 300
(Lee et al., 2018)
orthosilicate; HCl
XPS, PL

(𝑋𝑑𝑒𝑣𝑒𝑙𝑜𝑝𝑒𝑑 ‒ 𝑇𝑖𝑡𝑎𝑛𝑖𝑎 ‒ 𝑋𝑃25) ∗ 100


a 𝐸𝑛ℎ𝑎𝑛𝑐𝑒𝑚𝑒𝑛𝑡 𝑓𝑎𝑐𝑡𝑜𝑟 = where X is pollutant removal efficiency or degradation rate
𝑋𝑃25

Titanium butoxide (TBOT), TIP (Titanium isopropoxide), X-ray diffraction (XRD), Brunauer-Emmett-Teller (BET), Transmission electron microscopy (TEM), Diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS), Nitrogen adsorption-desorption (N2 ads-des),
Photoluminescence spectroscopy (PL), Anatase (A), Brookite (B), Rutile (R), Titanate (T), Scanning electron microscopy (SEM), Fourier transform infrared (FTIR), X-ray photoelectron spectroscopy (XPS), Thermal gravimetric analysis (TGA), Differential scanning calorimetry
(DSC), Energy dispersive X-ray (EDX), Electrochemical impedance spectroscopy (EIS), Temperature programmed desorption (TPD), Electron paramagnetic resonance (EPR), Electron spin resonance (ESR), Inductive coupled plasma-optical emission spectroscopy (ICP-OES),
Benzene, toluene, ethyl benzene, and o-xylene (BTEX)

11
ACCEPTED MANUSCRIPT

2.1. TiO2 nanoparticles and microspheres (0-D)

Zero-dimensional TiO2 photocatalysts are the most utilized architectures in photocatalytic air

purification. Considering the fact that photocatalytic reactions happen on the surface of

catalyst, it is ideal to maximize the exposure of photocatalyst to light and air stream in order

to improve the quantum efficiency and mass transfer rate. Due to the small primary particle

size, nanoparticles usually possess high specific surface area and high pore volume, which

can enhance contaminant adsorption, light-harvesting ability and photocatalytic

decomposition.

2.1.1. Nanoparticles

Anatase, rutile, and brookite are the main crystal phases of titanium dioxide, each possessing

distinct physical and chemical properties. Although there is limited work on photoactivity of

monophase brookite or rutile, the majority of previous studies demonstrated that anatase is

the most active polymorph of titania in gas-phase PCO. Various approaches have been

employed to control the phase composition of TiO2: pH of reaction solution, hydrothermal

preparation duration and temperature, type of solvent, post-synthesis calcination, etc.

Altering the pH of hydrothermal reaction medium (e.g. via addition of acid/base) is a facile

and efficient method to adjust the degree of crystallinity and crystal phase of TiO2. In this

regard, Van der Meulen et al. (van der Meulen et al., 2007) synthesized mixed-phase anatase-

rutile TiO2 through solvothermal treatment of microemulsions containing Triton X-100 and

n-hexanol as surfactant, cyclohexane as the oil phase, and HCl or HNO3 as the crystal phase

determining agent. It was found that HCl leads to formation of pure rutile structure while

pure anatase can be obtained with HNO3. This observation was attributed to the presence of

different ions during the synthesis (Cl- or NO3-), which can promote the formation of a

particular crystalline phase. Anatase, rutile, and mixed-phase anatase-rutile photocatalysts

12
ACCEPTED MANUSCRIPT

showed 15, 6, 9% propane removal efficiency, respectively. In another study, tricrystalline

(anatase, rutile, and brookite) titania was fabricated by hydrothermal treatment of amorphous

TiO2 in concentrated HNO3 aqueous solution at 180 ˚C for 24 h (Chen et al., 2015b). X-ray

diffraction (XRD) results indicated that as the acidity of reaction medium was lowered,

anatase and rutile contents respectively increased and decreased, while brookite formation

was minimally affected. Concomitantly, the degree of crystallinity, crystal size, and pore size

declined and the surface area improved. Toluene removal efficiency enhanced with acidity,

peaked at HNO3 to titanium butoxide (TBOT) molar ratio of one and upon further increment

in HNO3 concentration sharply diminished. The optimum photocatalytic performance toward

degradation of toluene (about 3.85-fold higher than that of P25) was achieved over titania

with 80.7%, 3.7%, and 15.6% of anatase, rutile, and brookite, respectively. The superiority of

the developed titania was in part attributed to its greater OH• formation rate, which was

evidenced via photoluminescence spectroscopy (PL). The heterojunctions formed between

different crystal phases facilitate the transfer of photogenerated electrons from one phase to

another and mitigate e--h+ recombination, which in turn boosts OH• generation. This

phenomenon was also previously examined via electron paramagnetic resonance (EPR)

method and attributed to electron transfer between phases (Hurum et al., 2005).

Similar relationships between phase content and acidity were also found by other researchers

during hydrothermal treatment of hydrous oxide in ethanol-water solution (Deng et al., 2002;

Wu et al., 2004; Cano-Casanova et al., 2018). Recently, Cano-Casanova et al. (Cano-

Casanova et al., 2018) conducted a detailed study on crystalline structure and surface OH

content of TiO2 photocatalysts prepared by solvothermal method using titanium isopropoxide

(TIP) and HCl (0.5-12 M). The concentration of HCl had a great impact on crystal phase and

size, OH density, and consequently the activity for propene oxidation. As the HCl

concentration raised (up to 3 M), the anatase phase content decreased while rutile and

13
ACCEPTED MANUSCRIPT

brookite contents steadily increased. At higher HCl concentrations, the anatase percentage

increased with solution acidity and naturally rutile and brookite contents declined. Based on

the results of thermal gravimetric analysis (TGA), it was determined that the amount of

surface OH groups decreases with HCl concentration. The optimum propene conversion

(50%) was obtained on titania sample with 0.8 M HCl, most probably because among all

samples it had the largest surface area, porosity and OH content, and the smallest crystal

sizes. Wu et al. (Wu et al., 2004) observed that with rising the concentration of HNO3 (from 0

to 1 M) the rutile phase content steadily increases from 0 to 100%. Surprisingly, rutile crystal

size, surface area and activity towards methanol/hexane did not show predictable trends with

acidity. The maximum hexane (25.5%) and methanol (45.2%) removal efficiencies were

achieved over titania samples with 12 and 15% rutile phase, respectively. Evidently, the

lower e--h+ recombination rate and higher crystallinity of biphasic titania are counteracted by

the loss of surface area at higher acidities, leading to appearance of activity optimums. Chen

et al. (Chen et al., 2015a) applied various acids (HF, HCl, HNO3, H2SO4, and CH3COOH)

during hydrothermal treatment to adjust the structural and crystalline properties of

photocatalysts. In the first step, titanate nanotubes (TiNTs) were prepared by hydrothermal

treatment of TiO2 anatase powder in NaOH solution at 180 ˚C for 12 h. Subsequently, TiNTs

endured acid-assisted hydrothermal reactions for 24 h at 180 ˚C. In the presence of HF,

H2SO4 or CH3COOH, anatase was the only detected polymorph while by employing HCl or

HNO3 tricrystalline titania could be obtained. Scanning electron microscopy (SEM) images

demonstrated that acid type can affect the morphology of final products: Octahedral-shaped

particles with HF, spherical and/or rod-like particles with HNO3 and HCl, and spherical

particles with H2SO4 and CH3COOH. The best photocatalytic activity towards toluene was

acquired on samples that were prepared in the presence of CH3COOH, though these samples

did not necessarily possess the largest surface areas or highest crystallinities. Pure anatase

14
ACCEPTED MANUSCRIPT

TiO2 synthesized in 15 M acetic acid exhibited the best toluene removal efficiency (94%),

more than twice that of P25 (44%). Based on Fourier transform infrared (FTIR) results, it was

suggested that surface modification by acetic acid could result in less hydrophilic TiO2

particles which can better adsorb toluene.

The solvent in hydrothermal/solvothermal synthesis can also exert influence on the key

characteristics of titania photocatalysts. In this regard, Shi and Weng (Shi and Weng, 2008)

have reported the synthesis of TiO2 with controllable anatase content by varying the volume

ratio of water to ethanol (as co-solvent) in a low temperature preparation route. XRD analysis

confirmed that increment in ethanol content in the reaction medium leads to higher anatase

fraction and smaller crystallite sizes (for both anatase and rutile), suggesting alcohol

promotes anatase formation and inhibits crystallite growth. It was noted that the

photocatalytic decomposition of formaldehyde gradually improves with raising anatase

content and reaches its optimum over the titania sample with 59% anatase and 41% rutile.

The antenna effect (by the rutile phase) and transfer of electrons from rutile to anatase via

formed junction between the two phases enhanced the generation of active radicals and PCO

reactions. In contrast, in a study conducted by Maira et al. (Maira et al., 2000), increasing the

concentration of alcohol (isopropanol) in the reactor led to larger primary particle sizes.

Transmission electron microscopy (TEM) images illustrated that the secondary particle size

decreases rapidly with water content during sol-gel synthesis, which can be ascribed to a

more homogeneous nucleation rate and smaller gel spheres. The authors reported that the

photocatalyst with the primary particle size of 7 nm and secondary particle size of 100 nm

achieved the best activity in PCO of trichloroethylene (TCE).

A few studies have investigated the structural differences between photocatalysts underwent

hydrothermal or thermal treatment (Maira et al., 2001; Wu et al., 2007; Ardizzone et al.,

2008).Wu et al. (Wu et al., 2007) fabricated different TiO2 nanoparticles via four preparation

15
ACCEPTED MANUSCRIPT

procedures and evaluated their performance for benzene degradation. First, mother-liquid was

prepared by dropwise addition of TBOT-ethanol solution to a water-ethanol solution under

continuous stirring. In the solvothermal route, the mother-liquid aged for 12 h at 200 ˚C in

autoclave, while in the thermal route, calcination at 450 ˚C for 3 h was performed. Evidenced

from TEM images, one main difference between these samples is that the crystallite size

distribution is broader for the latter owing to the heterogeneous condition during

crystallization. As displayed in Fig. 2a, the lattice images indicate the high crystallinity of

sample prepared by solvothermal treatment. On the contrary, TiO2 obtained from the thermal

route (Fig. 2b) has a zigzag surface and the development of crystallites is incomplete. During

calcination, the transform and aggregate of crystallite takes place before the crystallite

develops well. The photocatalytic activity of the TiO2 by solvothermal method in degradation

of 125.4 ppm benzene was 4.1 times higher than that prepared via calcination.

Fig. 2. High resolution TEM images and related electron diffraction patterns of samples
prepared via (a) solvothermal route and (b) thermal route (Wu et al., 2007).

Considering the crucial role of surface hydroxyl groups in the adsorption of pollutants and the

formation of hydroxyl radicals, the impact of preparation technique on the population of OH

groups on TiO2 surface is highly relevant to air purification (Mamaghani et al., 2018a). In this

16
ACCEPTED MANUSCRIPT

regard, some studies proposed that hydrothermal route can preserve the hydroxylation state of

TiO2 surface (Maira et al., 2001; Wu et al., 2005; Ardizzone et al., 2008). Ardizzone et al.

(Ardizzone et al., 2008) compared the population of OH groups (via X-ray photoelectron

spectroscopy (XPS)) on the surface of titania samples subjected to either thermal (5 h at 300

°C) or hydrothermal treatment (500 h at 60 °C). It was found that the concentration of surface

OH groups for hydrothermally-prepared TiO2 is roughly 33% higher than that of the calcined

one. Similar results were reported in a work by Maira et al. (Maira et al., 2001) in which two

anatase TiO2 photocatalysts were prepared from amorphous titania either via calcination at

450 ˚C or hydrothermal reactions at 200 ˚C. The hydrothermally-prepared sample exhibited

higher activity and better stability during photocatalytic degradation of toluene. This was

attributed to the greater number of H-bonded hydroxyls on the surface of hydrothermally-

prepared TiO2 and its ability to regenerate OH groups in the presence of water vapour. A

comparison between the FTIR spectra of titania before and after the experiments indicated

that the deactivation of titania prepared by calcination could be due to severe surface

dihydroxylation.

2.1.2. Micro and hollow spheres

Synthesis of porous TiO2 microspheres has been the focus of numerous researches owing to

their high surface-to-volume ratio, large surface area, and enhanced light harvesting ability

(Li and Li, 2001; Wang et al., 2013; Zheng et al., 2014). The most conventional technique for

fabrication of porous spheres is templating strategy in which removable or sacrificial

templates (e.g. polymer latex, carbon, silica, surfactants, and oil droplets) are used to achieve

desired spatial arrangements (Zhong et al., 2000; Caruso et al., 2001; Ren et al., 2003; Zhang

et al., 2004; Ismail and Bahnemann, 2011; Li et al., 2016). Obvious shortcomings of

templating routes such as the need for expensive templates and unwanted structural changes

during template removal have encouraged scientists to develop template-free methods for

17
ACCEPTED MANUSCRIPT

production of porous nanostructures (Collins et al., 2004; Yu et al., 2007c; Yu et al., 2007d;

Zhang et al., 2009). Most of these methods are based on metastable processes: Kirkendall

effect, Ostwald ripening, and chemically induced self-transformation (Yang and Zeng, 2004;

Yin et al., 2004; Yu et al., 2006d; Xiang et al., 2010).

Zhou et al. (Zhou et al., 2010) have reported the preparation of mesoporous TiO2

microspheres by hydrothermal treatment of Ti(SO4)2 in CO(NH2)2 aqueous solution at 160 ˚C

for various durations. As the hydrothermal time extended, the crystal size and relative anatase

crystallinity increased since a longer reaction time is beneficial to Ostwald ripening and

growth of nanocrystals. Simultaneously, the hysteresis loops moved to a higher relative

pressure and lower adsorbed volume region, implying increment in average pore size and

decrement in surface area. These result from the collapse of smaller pores prior to the bigger

ones and excessive crystal growth during hydrothermal treatment. The photocatalytic activity

enhanced with prolonging the hydrothermal time, reached its maximum at 7 h, and

considerably declined for the sample treated for 12 h. Acetone degradation rate on

mesoporous TiO2 spheres treated for 7 h was roughly 2.5 times higher than that of P25, due

to the large surface area, porous structure, and good crystallinity of the prepared titania. In a

similar work, porous TiO2 microspheres with distinct structural properties were synthesized

by varying the hydrothermal temperature and duration (Du et al., 2011). SEM images

confirmed that microspheres diameter and morphology irregularity increase with temperature.

High hydrothermal temperature leads to extensive nucleation at the early stages of

preparation, which limits TiO2 crystals adhesion to the surfaces of precursor in the bulk

solution, resulting in the formation of particles with irregular shapes. The photoactivity

enhanced with increasing the hydrothermal temperature from 140 to 180 ˚C and further

increment to 200 ˚C adversely affected the activity due to the destruction of microspheres

structure.

18
ACCEPTED MANUSCRIPT

Several groups have successfully fabricated TiO2 hollow spheres via hydrothermal method in

the presence of fluoride sources (Yu and Shi, 2010; Yu and Zhang, 2010; Liang et al., 2017) .

Yu and Zhang (Yu and Zhang, 2010) obtained TiO2 hollow spheres by solvothermal

treatment of amorphous solid spheres in NH4F solution. The TiO2 solid spheres were

prepared by hydrolysis of TBOT in a mixture of ethanol and KCl under stirring, followed by

24 h aging. Next, the as-prepared TiO2 spheres were mixed with varying amounts of NH4F

aqueous solution and maintained in autoclave for crystallization at 180 ˚C for 12 h. In the

absence of NH4F, anatase and brookite polymorphs existed in the crystalline structure,

whereas introducing a small amount of NH4F suppressed brookite formation and promoted

anatase crystal growth. SEM images showed that at higher NH4F concentrations, surface

roughness of hollow spheres and the number of broken spheres increase due to crystal growth

and reduction in the thickness of TiO2 shells. Solvothermal treatment with NH4Cl or NH4Br

(instead of NH4F) produced crystalline solid spheres, indicating F¯ ions promote or induce

the hollowing of TiO2 solid spheres. A formation mechanism was proposed based on the

structural and morphological properties of products obtained at different solvothermal times

(Fig. 3). Crystallization initiates at the surface of solid spheres due to the direct contact with

aqueous medium and continues until the surface layer and the medium reaches equilibrium.

Subsequently, due to the higher solubility and surface energy of amorphous core compared to

the outer surface, the dissolution (via F¯ ions) and hollowing processes preferentially act on

the centre of sphere rather than outer regions. Photocatalytic activity greatly varied with

NH4F concentration and exhibited an optimum, possibly due to the opposite trends in degree

of crystallinity and surface area with acidity. Acetone degradation rate constant reached its

optimum value at NH4F to TiO2 molar ratio of 1, exceeding that of P25 by a factor of 2.3.

Large surface area, good crystallinity, hollow and porous structure, and presence of F¯ ions

19
ACCEPTED MANUSCRIPT

on the surface were considered as the main contributors to the high activity of developed

samples in the presence of NH4F.

Fig. 3. Proposed formation mechanism of anatase TiO2 hollow spheres. The left and
right panels present the TEM images and XRD patterns of intermediate products at 180
˚C for 0-720 min of solvothermal treatment (Yu and Zhang, 2010).

Yu and Shi (Yu and Shi, 2010) have reported the preparation of TiO2 hollow microspheres

via a one-pot hydrothermal route using Ti(SO4)2 and trifluoroacetic acid (TFA) as titanium

and F¯ sources, respectively (Fig. 4). The molar ratio of TFA to Ti(SO4)2 in the reaction

medium were varied to control the key properties of titania powder (e.g. crystal phase, crystal

size, and surface area). At higher TFA concentrations, the degree of crystallinity and anatase

crystal size enhanced due to the presence of F¯ in the environment. N2 adsorption-desorption

20
ACCEPTED MANUSCRIPT

isotherms revealed existence of mesopores and macropores (bimodal pore size distributions),

which resulted from the intra-agglomerated primary particles and inter-aggregated secondary

particles, respectively. The authors attributed the formation of hollow structures to localized

Ostwald ripening and chemically induced self-transformation mechanisms, which were also

put forward by other researchers (Shengwei et al., 2009; Yu et al., 2009c). PL results showed

that the formation rate of OH• on samples prepared in the presence of TFA was much higher

than that of titania prepared in pure water. This is understandable considering that surface-

adsorbed TFA complexes and F¯ can trap photogenerated electrons, lower recombination

rate, and consequently boost OH• generation (Yu et al., 2003; Yu et al., 2009a). Titania

hollow microspheres substantially outperformed P25 in photocatalytic degradation of acetone

in air (rate constant of 0.064 vs 0.035 min-1). This was explained by underlining that the

hollow and bimodal porous structure assist light harvesting and mass transfer of reactants to

the active sites, while fluoride species prolong e--h+ lifetime.

Fig. 4. SEM (a) and TEM (b) images of TiO2 sample prepared at molar ratio of TFA to
Ti(SO4)2 equal to 2 (Yu and Shi, 2010).

Liang et al. (Liang et al., 2017) synthesized surface-fluorinated hollow microspheres with

excellent thermal stability via hydrothermal reaction of Ti(SO4)2-NH4HF-H2O2 solution at

21
ACCEPTED MANUSCRIPT

180 °C. The impact of calcination temperature (300-1100 °C) on the structure and

photoactivity of hollow microspheres in acetone degradation was examined. With rising the

calcination temperature from 300 to 1000 °C, anatase crystal size and crystallinity gradually

increased from 79.0 to 97.7 nm and from 1.00 to 1.53, respectively. Transformation of

anatase to rutile phase is seen around 600 °C; however for the fluorinated microspheres, the

XRD peaks associated with rutile phase first appeared after calcination at 1100 °C, showing

high thermal stability. A morphology evolution was proposed based on the XRD, SEM, and

TEM analyses of TiO2 samples prepared at different calcination temperatures (Fig. 5).

Hollow microspheres are composed of nano-size building blocks, which gradually evolves

with calcination temperature from truncated bipyramidal nanoparticles with exposed [001]

facets at 300 °C to solid nanospheres at 1100 °C. The photocatalytic activity of as-prepared

hollow spheres experienced a 40% decline upon calcination at 300 °C due to the desorption

of fluoride ions from the surface, which are known to be capable of forming highly reactive

free OH•. From 300 to 900 °C, the activity improved with calcination temperature due to

enhanced crystallinity. Further increment in temperature from 900 to 1100 °C led to a

performance drop caused by the reduction in surface area and appearance of rutile phase.

22
ACCEPTED MANUSCRIPT

Fig. 5. SEM and TEM images of TiO2 hollow microspheres prepared at different
calcination temperatures (Liang et al., 2017).
2.1.3. Nanoparticles-graphene composite

One of the key steps in photocatalytic air purification is the adsorption of pollutants onto the

catalyst surface. Due to the low concentrations of pollutants (ppb range) and short contact

times (milli seconds), adsorption efficiency greatly influences the overall performance of air

purifiers (Yao et al., 2017; Sboui et al., 2018). Although titanium dioxide nanoparticles offer

23
ACCEPTED MANUSCRIPT

high photoactivity, their low adsorption affinity towards air pollutants (e.g. VOCs) and small

surface area diminish the reaction rate and removal efficiency. Therefore, it is desired to

couple titanium dioxide with materials that could enhance the adsorption of pollutants under

realistic conditions (Kibanova et al., 2009; Kibanova et al., 2012). In this regard, various

carbonaceous materials (e.g. activated carbon, carbon nanotubes, carbon fibers, graphene

oxide (GO), and g-C3N4) have been investigated as supports. Among various materials,

graphene oxide has attracted much attention in the field of photocatalysis owed to its large

surface area, high electron mobility, hydrophobic surface, and high adsorption affinity for

organic pollutants through Π-Π interactions.

Yu et al. (Yu et al., 2018) evaluated the degradation rate of formaldehyde under visible light

over TiO2-reduced graphene oxide (rGO) prepared via hydrothermal reaction between TBOT

and graphite powder. Formaldehyde (initial concentration=500 ppb) removal efficiencies

over TiO2, rGO, and rGO/TiO2 after 100 min irradiation were 10, 20, and 60% respectively.

It was suggested that rGO can serve as an electron sink for electrons in TiO2 conduction band

and separate e--h+ pairs through interfacial charge transfer. In other words, rGO extends the

lifetime of charge carriers, which in turn increases the possibility of photochemical reactions

and formation of reactive radicals (e.g. OH• and O2•-).

In another study, rGO-TiO2 composite was synthesized by a simple hydrothermal method

using P25 and GO, and subsequently coated on polyacrylonirile (PAN) scaffolds via

electrospraying (Roso et al., 2015). The superior activity of rGO-TiO2 composite towards

gaseous methanol with respect to TiO2 was attributed to its enhanced charge transfer, larger

adsorption ability, and narrower band-gap. During the hydrothermal synthesis, unpaired Π

electrons of GO can bond with free electrons of TiO2 and form Ti-O-C structure, which

narrows the composite band-gap. Nevertheless, it is noteworthy that at excessive graphene

24
ACCEPTED MANUSCRIPT

loadings, the aforementioned positive effects of rGO were overpowered by the reduction in

composite light utilization efficiency and the removal efficiency declined.

Yadav and Kim (Yadav and Kim, 2016) prepared TiO2-rGO nanocomposites with different

GO loadings (0.25-2 wt %) via solvothermal method for photocatalytic oxidation of benzene

in air. The SEM image and EDS elemental mapping of TiO2-rGO photocatalyst (Fig. 6)

exhibited that anatase TiO2 particles are dispersed on the surface of GO sheet. Expectedly,

UV-vis spectroscopy showed higher absorption in the visible light range for all TiO2-rGO

composites compared to pure TiO2, meaning decrement in TiO2 band-gap. The hypothesis

that coupling titania with rGO can alleviate the recombination of charge carriers was

supported by the results of photoluminescence spectroscopy, which showed much lower PL

intensity for TiO2-rGO compared to TiO2. TiO2-rGO composite with 0.25% GO loading

exhibited the highest activity and reached 100% benzene removal (vs 67% for TiO2) within

50 min. Nevertheless, beyond 0.25 wt%, the photocatalytic activity demonstrated a

downward trend with GO loading. It was suggested that after a certain GO concentration,

graphene oxide itself can act as a recombination center for e--h+ pairs and also strongly blocks

the UV light.

25
ACCEPTED MANUSCRIPT

Fig. 6. SEM image of TiO2-graphene oxide (0.25 wt%) and the corresponding EDS
elemental mapping (Yadav and Kim, 2016).

Ebrahimi and Fatemi (Ebrahimi and Fatemi, 2017) investigated the PCO of high

concentration of acetaldehyde (500 ppm) over P25-rGO at two different GO contents (0.5 and

2.5 wt%). Graphene loading had small effect on surface area and crystallinity, while it greatly

influenced the surface hydrophilicity and band-gap energy. Comparing the FTIR spectra of

P25 and P25-rGO revealed that the 1630 and 3430 cm-1 peaks, which are associated with

adsorbed water and hydroxyls groups, are more intense for P25, indicating P25 surface is

more hydrophilic. On the other hand, by introducing rGO into titania structure the 3.02 eV

band gap of P25 was reduced to 2.85 eV, enabling P25-rGO excitation under visible light and

efficient generation of e--h+ pairs. Acetaldehyde removal efficiencies for P25, P25-rGO-0.5%

and P25-rGO-2.5% were 83, 91, and 63%, respectively after 10 min UV light irradiation. The

difference between P25 and P25-rGO-0.5% removal efficiencies under visible light (36 vs.

26
ACCEPTED MANUSCRIPT

55%) was more noticeable due to the narrower band gap and lower e--h+ recombination for

the nanocomposite.

In a detailed and interesting work, TiO2-graphene (0-30 wt%) nanocomposites were

fabricated by solvothermal treatment of GO and P25 in ethanol-water solution and tested for

photocatalytic degradation of benzene (Zhang et al., 2010). Incorporation of graphene had a

tremendous effect on the optical properties of P25 and led to a red shift in the absorption

edge. TiO2-graphene band gap steadily decreased with increasing the amount of graphene in

the composite, from 3.36 eV for bare TiO2 to 2.83 eV for TiO2-graphene-30%. All TiO2-

graphene photocatalysts outperformed P25 in terms of removal efficiency, pollutant

mineralization to CO2, and durability. In particular, P25-graphene-0.5% nanocomposite

showed the highest benzene removal efficiency under UV light (6.4% after 28h of operation),

which was roughly 5 times higher than that of P25. The authors ascribed the superior

performance of P25-graphene to improved adsorptivity of benzene, higher light absorption,

and longer lifetime of charge carriers which was confirmed via electron spin resonance

(ESR). ESR spectrum of P25-graphene exhibited much stronger signals for OH• and O2•- in

comparison to those of P25. It was suggested that the interactions between hydrophilic

functional groups on graphene oxide and surface hydroxyl groups on titania leads to

dispersion of TiO2 nanoparticles over graphene sheets. This close contact between the two

media and the fact that graphene is an excellent electron acceptor can facilitate interfacial

charge transfer. On the other hand, the conduction band position of anatase is -0.24 V and the

potential of graphene/graphene•- is -0.08 V (vs. SHE); thus theoretically, upon illumination,

electrons can travel from the conduction band of TiO2 to the graphene sheets (Wang et al.,

2012).

Contrary to the previous works, Suave et al. (Suave et al., 2017) reported that coupling TiO2

with graphene, even at very low rGO content of 1%, adversely affected the photocatalytic

27
ACCEPTED MANUSCRIPT

activity. Nitrogen oxide conversion over pure titania nanoparticles was 35%, while for the

TiO2-rGO composite it decreased to 25%. They suggested that the preparation procedure of

the composite could not result in a close contact between TiO2 nanoparticles and graphene

oxide and, therefore, the phenomenon of interfacial charge transfer did not take place

efficiently.

Li et al. (Li et al., 2017) employed a facile solvothermal route to prepare g-C3N4/rutile-

brookite TiO2-xNy composites using TiCl3, hexamethylenetetramine, and melamine as Ti, N,

and g-C3N4 source respectively. The yielded composites showed significant absorption in

visible light region due to the incorporation of g-C3N4 into TiO2 structure and N-doping.

Photoluminescence and photocurrent measurements clearly indicated that the magnitude of e-

-h+ recombination for g-C3N4/rutile-brookite TiO2-xNy composites is much smaller than those

of g-C3N4 and TiO2. This observation was attributed to the formation of a Z-scheme

photocatalytic system between g-C3N4 and TiO2-xNy under hydrothermal conditions, which

allows efficient charge transfer between semiconductors. N-doped brookite-rutile

photocatalysts (i.e. without g-C3N4) reached 60 and 72% NO removal efficiency under visible

and UV light, respectively, which was a great improvement over P25 (30% for visible and

50% for UV light). Surprisingly, adding g-C3N4 to the N-doped TiO2 had different impacts on

the photocatalytic performance depending on the crystalline phase of titania: promoting for

biphasic rutile-brookite and inhibiting for single brookite titania. It was proposed that the

electron transfer from brookite TiO2 to g-C3N4 is not as efficient as it is in g-C3N4/rutile-

brookite.

2.1.4. Nanoparticles/microspheres doping and surface modification

To date, numerous studies have been focused on boosting TiO2 photocatalytic activity by

deposition of noble metals (e.g., Au, Ag, Pt, Pd, silver), non-metals doping (e.g. N and C),

and surface modification (Zhang et al., 2014; Dong et al., 2018). It is generally accepted that

28
ACCEPTED MANUSCRIPT

the deposited metal acts as a reservoir for photo-induced electrons and reduces charge

recombination in semiconductors (Fig. 7) (Subramanian et al., 2004; Wu et al., 2009). On the

other hand, noble metals deposition can shift the photoresponse of titania into the visible

region due to surface plasmon resonance (SPR) (Jain and El-Sayed, 2007; Shi et al., 2017).

Moreover, formation of a Schottky barrier at the metal-semiconductor interface facilitates the

interfacial electron transfer and promotes charge carrier separation (Fig. 7b) (Vijayan et al.,

2010). Nevertheless, it is worth mentioning that unsuccessful doping could create unwanted

recombination centres or localized band gap states, which prevent efficient separation of e--h+

pairs and/or place the conduction band edge below the O2 affinity level (Mattsson et al.,

2006).

Fig. 7. (a) Transfer and separation of photogenerated electrons and holes and (b)
Schottky barrier at the metal-semiconductor interface (Yu et al., 2009b).

Yu et al. (Yu et al., 2009b) prepared Au doped TiO2 microspheres by solvothermal treatment

of precipitates of TBOT in a solution consisting of water, ethanol, and Au colloid particles.

The formation of microspheres was believed to be a result of thermodynamically driven self-

aggregation of countless metastable tiny nanoparticles. XRD results indicated that the anatase

crystallite size declines with increasing Au content in the reaction medium, since Au particles

suppress the growth of anatase crystals by providing dissimilar boundaries and limiting mass

transportation. The migration of electrons to Au clusters and the resulting enhancement in

29
ACCEPTED MANUSCRIPT

separation of charge carriers upon doping was also confirmed by PL spectra (see Fig. 8). As

RAu (atomic percentage ratio of Au to Ti) is raised from 0 to 0.00425, a noticeable decrement

in PL intensity can be seen, which implies low recombination rate. Nevertheless, at a higher

RAu value of 0.017, PL peaks became stronger due to the fact that excess Au atoms on TiO2

act as recombination centres. It is interesting to note that the photocatalytic activity of Au–

TiO2 nanocomposites toward formaldehyde (inset in Fig. 8) agree well with the PL

intensities, emphasizing the crucial impact of charge carriers separation.

Fig. 8. PL spectra and reaction rate constants (inset) of P25 and Au-TiO2 microspheres
obtained at RAu=0 (a), 0.00085 (b), 0.00425 (c), and 0.017 (d); RAu is atomic percentage
ratio of Au to Ti (Yu et al., 2009b).

In another study, surface of hydrothermally prepared porous TiO2 microspheres were

modified with noble metal nanoparticles (Ag, Au, Pt, or Pd) using photodeposition method

(Grabowska et al., 2016). All noble metal doped microspheres exhibited high absorption in

the visible region (especially in the case of Au and Pd) which was attributed to the SPR of

noble metal particles. However, it is noteworthy that doping beyond optimum values brought

30
ACCEPTED MANUSCRIPT

about decrement in toluene removal efficiency, possibly because of the coverage of TiO2

active sites by metal nanoparticles and growth in the number of recombination sites. Among

the decorated microspheres with Ag, Au, Pt, or Pd, 1% Ag-TiO2 and 0.1% Ag-TiO2

microspheres showed the best activity under UV and visible irradiation, respectively.

Recently, Zeng et al. (Zeng et al., 2018) could synthesize TiO2-nanoparticles/Ag-nanowires

photocatalysts via solvothermal treatment of Ag nanowires and TBOT in ethanol-water

solution. Fig. 9 depicts the SEM and TEM images along with the EDS elemental mapping for

a single Ag nanowire coated with TiO2 nanoparticles. The 1-D Ag nanowire has a diameter of

40-50 nm (Fig. 9b), which increases to roughly 70 nm after the deposition of TiO2. The Ag,

Ti, and O elemental mapping (Fig. 9b) confirmed a uniform distribution for titania

nanoparticles along the nanowire under solvothermal environment. The HRTEM images (Fig.

9 c and d) show a clear overlap of the lattice boundaries between Ag and TiO2, proving a

close contact between the two phases. The removal efficiency of acetaldehyde over TiO2/Ag

nanowires composite was 72%, almost twice that of pure TiO2 nanoparticles. The outstanding

performance of the composite was ascribed to the fact that Ag nanowires can not only

increase the light absorption (through SPR effect), but also serve as reservoir for electrons in

TiO2, allowing a better separation of charge carriers.

31
ACCEPTED MANUSCRIPT

Fig. 9. (a) SEM, (b) EDS elemental mapping images of Ag, Ti and O, and (c, d) HRTEM
images of the interface between Ag nanowires and TiO2 nanoparticles (Zeng et al., 2018)

Corrosion or dissolution of noble metal particles during PCO is an issue when metal

nanocrystals are anchored on the surfaces of the semiconductors and exposed to airflow.

Hetero-structures with metal core and semiconductor shell have been proposed to resolve this

issue (Subramanian et al., 2001; Kamat, 2007). Core-shell Au-TiO2 nanoparticles with radial

wedge-shaped antennae were prepared by hydrothermal route and the performance was

evaluated in PCO of acetaldehyde (Wu et al., 2009). Based on TEM images obtained at each

preparation step, Wu et al. proposed a morphological evolution mechanism for Au-TiO2

nanoparticles (Fig. 10). It was showed that both in terms of removal efficiency and

mineralization, the developed catalyst outperformed P25. For instance, after 15 min under

UV light, core-shell catalysts achieved 63% acetaldehyde removal while P25 barely reached

8%. This difference is even more noticeable under visible light source, where after 210 min

32
ACCEPTED MANUSCRIPT

only 6% degradation occurred on P25 versus 80% on the prepared core-shell Au-TiO2

nanoparticles.

Fig. 10. Formation process of core-shell Au-TiO2 nanoparticles with truncated wedge-
shaped morphology (Wu et al., 2009).

Regarding the non-metal dopants, Zhao et al. (Zhao et al., 2012) reported that nitrogen

doping of P25 not only could boost toluene removal efficiency but also substantially reduced

the health risks associated with by-products. The central reason for the noted enhancement in

photoactivity was believed to be the more efficient separation of charge carriers on the N-

doped sample. The stronger OH• and O2•- peaks in ESR spectra of N-doped titania with

respect to P25 along with higher photocurrent density under UV irradiation for the doped

sample confirmed presence of a larger number of electrons and holes. In another work, Shin

et al. (Shin et al., 2015) prepared N-F doped TiO2 by solvothermal method, using TBOT,

urea, and ammonium fluoride as Ti, N, and F sources, respectively, and ethanol as the

reaction medium. After doping titania with nitrogen and fluoride, three strong peaks appeared

in the XPS spectra which were assigned to F‾ ions adsorbed on the surface of titania, F atoms

that substituted O sites within the lattice, and molecularly chemisorbed N atoms. The

photocatalytic degradation efficiencies of toluene over N-F-TiO2, N-TiO2, and TiO2 under

visible light were 29.1, 17.4, and 15.7%, respectively. On the other hand, toluene removal

33
ACCEPTED MANUSCRIPT

efficiency improved with increasing N:F ratio from 3 to 6 and then gradually worsened as the

N:F ratio was raised to 16. This trend did not agree with the visible light absorption ability of

the samples (which constantly declined with N:F ratio) and instead assigned to the synergetic

effect of N and F.

Surface-fluorinated TiO2 materials have been suggested to be superior to their pure

counterparts owed to generation of mobile/free OH radicals (Park and Choi, 2004; Choi,

2006), though there are discrepancies. For instance, spin-trapping EPR measurements

validated that the concentration of hydroxyl radicals on fluorinated titanium dioxide under

UV irradiation is higher than that of unfluorinated one (Mrowetz and Selli, 2005, 2006) while

Lewandowski and Ollis (Lewandowski and Ollis, 2003) reported that the PCOs of aromatic

air pollutants were unaffected or even inhibited by surface fluorination of TiO2. In another

work, Mesoporous surface-fluorinated titania (F-TiO2) nanoparticles were synthesized in the

presence of NH4HF2 as F source via a one-step solvothermal method (Shengwei et al., 2009).

Owed to large surface area, mesoporous structure, small crystallite size and high crystallinity,

activity of all fluorinated TiO2 powders was higher than that of unmodified TiO2 and P25.

The photocatalytic activity of the best fluorinated TiO2 sample (prepared at atomic ratio of

fluorine to titanium of 0.5) towards acetone exceeded that of P25 by a factor of 3.

2.2. TiO2 nanotubes, nano-fibers, and nanowires (1-D)

TiO2 nanotubes (TNTs) and nanofibers, as one dimensional structures, offer a number of

characteristics which grant them excellent photocatalytic activity. These architectures not

only benefit from high surface-to-volume ratio and multiple pathways for diffusion of

molecules, but also have a low e--h+ recombination rate due to the fast interfacial charge

transfer (Yu et al., 2005; Toledo-Antonio et al., 2007; Liu et al., 2008; Zhao et al., 2010).

TNTs possess large amounts of hydroxyl functional groups on the interlayer region of their

walls due to their particular topography (Toledo Antonio et al., 2010; Rendón-Rivera et al.,

34
ACCEPTED MANUSCRIPT

2011). On the other hand, considering the high ion-exchange capacity of titanates, nanotubes

can be easily functionalized for instance to shift the photoresponse of titania into the visible

region (Sun and Li, 2003; Xu et al., 2012). As one of the pioneers, Kasuga et al. (Kasuga et

al., 1998) successfully fabricated Ti-based nanotubular structures via a facile hydrothermal

route through the transformation of TiO2 nanoparticles into nanotubes. Since then, many

groups developed TiO2 nanotubes by hydrothermal or thermal post treatment techniques

(Chen et al., 2002; Du et al., 2003; Yuan et al., 2004; Yu et al., 2006b). Most studies

synthesized titania nanotubes by hydrothermal treatment of P25 or anatase/rutile powders in

concentrated solution of NaOH or KOH (Shibata et al., 2007; Hernández-Alonso et al., 2011;

Nguyen and Bai, 2014).

Nguyen and Bai (Nguyen and Bai, 2014) prepared titania nanotubes by hydrothermal

treatment of P25 in a 10 M NaOH solution at 135 ˚C for 24 h. Subsequently, the solid

products were sonicated in HNO3 solution, filtered, washed with deionized water, dried at

120 ˚C, and finally calcinated at 400-600 ˚C for 2 h in air. Once the calcination temperature

exceeded 300 ˚C, the tubular structure started to transform to rod like (at 400 ˚C) and

particle-like (at 600 ˚C) structures (Fig. 11). Despite the drastic surface area loss and

excessive crystal growth at higher calcination temperatures, the sample calcinated at 500 ˚C

showed the best removal efficiency for NO and NO2, and surpassed those of P25 by a factor

of 2.85 and 2, respectively.

35
ACCEPTED MANUSCRIPT

Fig. 11. TEM results of sample calcinated at (a) 120 ˚C, (b) 400 ˚C, (c) 500 ˚C, and (d)
600 ˚C (Nguyen and Bai, 2014).

Hernández-Alonso et al. (Hernández-Alonso et al., 2011) demonstrated that mere

hydrothermal treatment in NaOH cannot assure formation of nanotubular structure and a

subsequent acid washing process (with HCl or HNO3) is essential. On the ground of TEM

images, they also highlighted that a 24 h hydrothermal treatment at 130 ˚C is not enough for

complete transformation of titania nanoparticles into titanate nanotubes and the resulted

products were branched nanostructures and/or poorly shaped nanotubes (see Fig. 12). By

optimizing the preparation parameters (e.g. acid concentration, TiO2 precursor, and

calcination temperature), some of the developed TNTs could reach high photocatalytic

activities for TCE degradation (up to 100%), which were comparable to that of P25.

36
ACCEPTED MANUSCRIPT

Fig. 12. TEM micrographs of nanostructures obtained from hydrothermal treatment of


P25 at 130 ˚C for (a) 24 h and (b) 72 h (Hernández-Alonso et al., 2011).

Inline with previous works on fabrication of 1-D nanostructures, Yu et al. (Yu et al., 2007a)

prepared hydrogen titanate nanowires (H2Ti6O13 or H2Ti8O17) via hydrothermal reaction of

P25 in KOH basic solution followed by calcination at high temperatures (300-900 ˚C). H-

titanate nanowires calcinated at 300 ˚C showed insignificant photocatalytic activity toward

acetone, which primarily stems from absence or low crystallization of anatase phase. With

increasing the calcination temperature, XRD diffraction peaks of anatase, brookite, and rutile

appeared and gradually strengthened, while surface area and porosity dramatically

diminished. The morphology changed from wire-like structure (with diameters < 5-10 nm) at

300 ˚C to nanorods with large diameters (> 200 nm) at 900 ˚C. The following factors were

mentioned to justify the superior activity of the sample calcinated at 500 ˚C: i) high surface

area and porosity, ii) co-presence of anatase and brookite (thus better separation of e--h+), and

iii) large band gap of brookite phase and more powerful redox ability.

As an alternative for high temperature calcination for transformation of titanate to anatase

phase, hydrothermal treatment at 200 °C for varying durations (1-24 h) in pure water was

applied (Yu et al., 2006c). Based on XRD analyses, after only 1 h, destruction of titanate

37
ACCEPTED MANUSCRIPT

structure and formation of anatase started and further increment in hydrothermal time entailed

enhancement of crystallization and crystallite size growth (from 8.1 nm after 1 h to 27.3 nm

after 24 h). On the other hand, by hydrothermal post-treatment of titanate nanotubes, the pore

size distribution became narrower and large pores (> 30 nm) almost disappeared due to the

collapse of the tubular structure and crystallization. Anatase TiO2 nanofibres that endured 7 h

of hydrothermal post-treatment presented the best photoactivity towards acetone with k value

of 5.18×10 -3 min-1 which was considerably higher than that of P25 (3.47×10 -3 min-1).

Chatterjee et al. (Chatterjee et al., 2010) have also succeeded in synthesizing one-dimensional

titanate nanostructures (trititanate and TiO2-B (a low-density titania polymorph)) via a simple

hydrothermal procedure at different temperatures. FTIR spectra of the sample treated at 100

°C exhibited much stronger absorbance intensity in the OH spectral region, 3800 to 2800 cm-
1, compared to that of the sample treated at 200 °C. The large concentration of OH groups on

the surface and the large surface area were accounted responsible for high degradation rate of

ethylene.

Similar to 0-D structures, doping or coupling with porous support materials can also be

applied to improve the photocatalytic activity of 1-D structures. In this regard, Vijayan et al.

(Vijayan et al., 2010) prepared platinum-doped titania nanotubes to achieve visible light

responsive photocatalysts. The dopant quantity exerted a great influence on the

morphological and electrical properties of the yielded titania samples, most significantly

surface area and band gap. SEM images revealed that the tubular morphology of titania is

well maintained up to a doping concentration of 1 mol % Pt and beyond that value nanotubes

gradually transform into nanoparticles. In addition, the band gap was narrowed from 3.16 eV

for undoped titania to 2.64 eV for 4% Pt-doped titania which showed enhanced absorption in

the 300-700 nm region. Unlike the undoped titania nanotubes and P25 which showed very

38
ACCEPTED MANUSCRIPT

small activity for PCO of acetaldehyde (about 10% removal), the 0.5% Pt-doped titania

nanotubes achieved around 45% acetaldehyde removal efficiency under visible radiation.

In another research, nitrogen-doped titania nanorodes with different N:Ti ratios were

synthesized and tested under visible light for benzene, toluene, ethyl benzene, o-xylene

(BTEX) photocatalytic degradation (Jo et al., 2014b). For all N:Ti ratios, the BTEX removal

efficiencies over the N-doped titania (between 19-90%) exceeded those obtained by pure

titania (between 4-19%) due to the fact that nitrogen atoms incorporation into TiO2 lattice

shifts the band-gap to the visible-light region. The activity of N-doped titania photocatalysts

increased with nitrogen doping up to N:Ti=0.5 and then followed a downward trajectory.

This can be explained from two perspectives: (i) the surface areas of N-doped samples

matched the above-mentioned trend and (ii) after a certain N:Ti ratio, the extra nitrogen

atoms in TiO2 may act as recombination centres for charge carriers, lowering the e--h+

lifetime.

Regarding the application of support materials, titania nanotubes were grown on carbon

fibers through hydrothermal treatment of TBOT-coated carbon sheets in concentrated NaOH

aqueous solution at 160 °C for 48 h (Jo et al., 2014a). TEM images ensured successful

formation of nanotubes (length and outer diameter of 20–110 and 6–11 nm, respectively) on

carbon fibers. The photocatalytic activity of TiO2 nanotubes-carbon fibers was compared

with that of TiO2 nanoparticles-carbon fibers prepared via a coating-calcination process.

Benzene, toluene, ethyl benzene, and o-xylene degradation efficiencies over TiO2 nanotubes-

carbon fibers were 81, 97, 99, and 99%, respectively, while those for TiO2 nanoparticles-

carbon fibers were 14, 42, 52, and 79%, respectively. The high adsorption capacity of carbon

fibers for aromatics compounds and the lower recombination rate of e--h+ pairs in one-

dimensional nanostructures were considered the main contributing factors to TiO2 nanotubes-

carbon fibers superior performance.

39
ACCEPTED MANUSCRIPT

Shi et al. (Shi et al., 2017) employed a combination of wet coating, hydrothermal growth, and

photoreduction techniques to prepare a novel ternary composite: Au-doped TiO2-nanowires

coated on porous carbon fibers (Au/TiO2NWs@CF). SEM images clearly showed that the

carbon fibers are almost completely encapsulated by the micrometer long Au-doped titania

nanowires. The band gap values for TiO2 NWs, Au/TiO2 NWs, TiO2NWs@CF, and

Au/TiO2NWs@CF were 3.07, 2.94, 2.78, and 2.56 eV, respectively. Given its smaller band

gap and augmented absorption in visible region (results from SPR of Au nanoparticles)

Au/TiO2NWs@CF can be easily excited by visible light. This along with the possibility of

electron transfer from TiO2 to both Au nanoparticles and carbon fibers drastically increased

the number of available e--h+ pairs for photocatalytic reactions. Styrene removal efficiencies

under visible light over TiO2 NWs, TiO2NWs@CF, Au/TiO2 NWs, and Au/TiO2NWs@CF

were 25, 40, 73, and 91%, respectively. The superior performance of ternary composite stems

from higher adsorption capacity, more efficient harvest of visible light, and lower

recombination rate of charge carriers.

2.3. TiO2 nanosheets (2-D)

Nanosheets are nano-sized flake-shape materials with a flat surface, small thickness (1-10

nm) and cross-sectional dimensions up to several tens of micrometers. Titania nanosheets can

be prepared via alkaline hydrothermal treatment of TiO2 powder or from protonic titanate

hydrates, followed by either calcination or HST reactions for crystallization (Nakata and

Fujishima, 2012; Verbruggen, 2015). For instance, Shibata et al. (Shibata et al., 2007)

synthesized TiO2 nanosheets via an alkaline hydrothermal process using protonic titanate

hydrates as precursor and evaluated the photocatalytic performance of nanosheets for

degradation of 2-propanol. They observed a decline in PCO of 2-propanol with increasing the

number of nanosheet layers, indicating that the photogenerated charge carriers in the inner

layers cannot easily diffuse to the surface.

40
ACCEPTED MANUSCRIPT

For 2-D structures, one of the promising strategies for boosting the photocatalytic

performance that has attracted tremendous attention is the usage of high-energy facets (Yang

et al., 2009; Wu et al., 2014). It is widely accepted that for anatase TiO2, (001) facets are

more reactive than the thermodynamically more stable (101) facets (Yang et al., 2009; Liu et

al., 2010b; Xiang et al., 2010). The surface energy of different facets of anatase TiO2 follows

the sequence of (110) 1.09 J.m-2> (001) 0.9 J.m-2> (100) 0.53 J.m-2> (101) 0.44 J.m-2

(Lazzeri et al., 2001). Considering this fact, many experimental works have focused on the

preparation of anatase crystals with high percentage of exposed (001) facets (Taguchi et al.,

2003; Sofianou et al., 2012; Wu et al., 2014).

Wu et al. (Wu et al., 2014) reported the preparation of surface-fluorinated TiO2 nanosheets

(F–TiO2) by hydrothermal method (190 ˚C for 24 h) employing hydrofluoric acid (HF) as

capping agent and TBOT as titanium source. The formation of (001) facets is facilitated by

HF, since F¯ adsorption on the surface lowers (001) facets free energy and makes them more

stable. As a result of the fluoride addition, crystallization and crystal size were enhanced

compared to the sample prepared in similar conditions except without using HF. SEM and

TEM images clearly revealed that pure TiO2 has a regular shape with mainly (101) facets

exposed, while F–TiO2 consists of sheet-shaped structures (side length of 30 nm) with 49.3 %

of the exposed facets being the highly active (001). Taking into account that in the XPS

spectra of F–TiO2 only fluoride peak was seen (i.e. no peak for F¯ ions in the lattice), the

authors suggested that the formation of surface fluoride (≡Ti-F) is due to a ligand exchange

reaction between F¯ and surface OH groups. The F-TiO2 catalyst showed good activity in

gaseous NH3 removal (around 60%), which was roughly twice that on P25. The superior

performance of the developed photocatalyst was mainly ascribed to ≡Ti-F groups on the

surface that hinder e--h+ recombination and the more active (001) facets. Surface ≡Ti-F

groups can act as electron-trapping sites owed to the strong electro-negativity of fluorine and

41
ACCEPTED MANUSCRIPT

transfer photogenerated electrons to O2 adsorbed on titania surface (Shengwei et al., 2009;

Yang et al., 2009; Yu et al., 2009b; Xiang et al., 2010).

Xiang et al. (Xiang et al., 2010) investigated the influence of HF concentration during

hydrothermal treatment on the morphology and photocatalytic activity of anatase nanosheets

in decomposition of acetone. Estimated from TEM results, in the absence of HF, the

synthesized titania consists mostly of aggregated nanoparticles with small percentage of

exposed (001) facets (< 10±1%). Comparatively, by increasing the concentration of HF in the

medium, the percentage of exposed (001) facets in TiO2 nanosheets and anatase crystallinity

experienced steady upward trends, while surface area declined. PL spectra obtained from

samples prepared at different HF concentrations suggest that up to a certain concentration, F¯

can markedly reduce the e--h+ recombination rate and beyond that optimum concentration F¯

ions themselves become recombination centres. Fluorinated TiO2 nanosheet with 71±7 %

exposed (001) facets had the best photocatalytic activity, and its acetone decomposition rate

was roughly nine times higher than that of P25. It has been postulated that another reason for

superior performance of fluorinated TiO2 is its ability to generate free OH• which has larger

redox potential compared to surface-adsorbed OH• (Minero et al., 2000; Pan et al., 2008; Yu

and Shi, 2010) (reactions (5) and (6)).


+ ∙ +
≡ Ti ‒ OH + ℎ 𝑣𝑏 ↔ ≡ 𝑇𝑖 ‒ 𝑂𝐻 (5)
‒ + ∙ +
≡ Ti ‒ F + 𝑂𝐻 + ℎ 𝑣𝑏 → ≡ 𝑇𝑖 ‒ 𝐹 + 𝑂𝐻𝑓𝑟𝑒𝑒 + 𝐻 (6)

An interesting observation in this study (Xiang et al., 2010) was that the photocatalyst with

the highest percentage of exposed (001) facets did not possess the best performance.

Consistently, in another work by Liu et al. (Liu et al., 2013), it was witnessed that

photocatalytic activity declines sharply when (001) facets fraction goes beyond 0.7. A

possible explanation for the noted behavior could be the transport directions of

photogenerated charge carriers, as depicted in Fig. 13. Many investigations have pointed to

42
ACCEPTED MANUSCRIPT

the fact that photogenerated holes and electrons preferentially migrate toward (001) and (101)

facets, respectively (Liu et al., 2011; Tachikawa et al., 2011; Roy et al., 2013). Consequently,

OH• or oxidation sites are mainly located on (001) facets while O2•- or reduction sites are

generated on (101) facets. Keeping this in mind, one can conclude that the lack of enough

(101) facets impedes separation of charge carriers and lowers the photocatalytic activity.

Fig. 13. (a) Schematic illustration of the spatial separation of h+ and e- on TiO2
nanosheets and (b) surface heterojunction between (001) and (101) facets (Yu et al.,
2014).

In another research, Sofianou et al. (2012) applied two different preparation routes to obtain

anatase nanoplates with dominant (001) facets and hollow microspheres made of anatase

nanocrystals. To obtain the former, a mixture of TIP, ethanol, and HF was transferred to the

autoclave and treated for 24 h at 180 ˚C. In the anatase hollow microspheres preparation

route, they solely changed the reaction medium to water and kept the other parameters

unchanged. The formation of TiO2 hollow microspheres was attributed to the fluoride-

mediated self-transformation of amorphous titania solid microspheres, in which fluoride

causes the dissolution of particles interior followed by mass transfer from the core to the

external surface. It is noteworthy that ethanol as the medium during solvothermal reactions

has the ability to control the hydrolysis process of titanium alkoxide and brings about

formation of nanoplates; while in pure water hydrolysis occurs fast and creates amorphous

43
ACCEPTED MANUSCRIPT

solid spheres. NO removal efficiency over nanoplates and hollow spheres were 14 and 7% ,

respectively, mainly because of the larger surface area and more (001) facets in nanoplates.

To shed light on the role of TiO2 facet in the adsorption process, Wang et al. (Wang et al.,

2015) prepared two titania samples with dominant (001) or (101) facets via hydrothermal

method (Fig. 14) and examined toluene adsorption using diffuse reflectance infrared Fourier

transform spectroscopy (DRIFTS). TiO2 nanosheet with 67% exposed (001) facets was

synthesized by mixing TBOT and 40% HF solution, followed by hydrothermal treatment at

180 ˚C for 24 h. For the preparation of TiO2 with 95% exposed (101) facets, a two-step

procedure was used: (i) P25 was hydrothermally treated in 10 M KOH solution at 200 ˚C for

48 h, and (ii) the synthesized sample in (i) was mixed with deionized water and treated at 200

˚C for 3 h. The higher adsoprtion capacity of TiO2 with dominant (001) facets for toluene was

attributed to the larger number of unsaturated five-coordinated Ti atoms (5c-Ti) in this

sample. DRIFTS spectra of pure titania samples demonstrated a stronger absorbance at 3449

cm-1 (assinged to terminal hydroxyl groups) for the sample with (001) facets. Based on this,

they proposed that unsaturated 5c-Ti sites take part in the formation of terminal Ti-OH

species, which are well-known active adsorption sites for VOC molecules (Bianchi et al.,

2014). The TiO2 nanosheet (Fig. 14c) significantly outperformed bipyramidal particles with

exposed (101) facets (Fig. 14f) in photocatalytic degradation of toluene: 99 versus 85%

toluene removal efficiency.

44
ACCEPTED MANUSCRIPT

Fig. 14. SEM images (a and d), TEM images (b and e), and schematic illustrations of
prepared TiO2 with (001) facets (a, b, and c) and TiO2 with (101) facets (d, e, and f)
(Wang et al., 2015).

Besides its ability to improve crystallinity and facilitate preparation of titania with dominant

(001) facet, fluoride can also increase the thermal stability of anatase titania by shifting the

anatase-to-rutile transformation temperature to higher values (Lv et al., 2009). Lv and co-

workers (Lv et al., 2011) fabricated surface-fluorinated anatase TiO2 nanosheets with

dominant (001) facets and high crystal phase transition temperature (up to 1100 ˚C). In a

typical preparation procedure, TBOT and HF 40% solution was mixed and hydrothermally

treated in autoclave at 200 ˚C for 24 h, followed by calcination at temperatures ranging from

300 to 1250 ˚C for 2 h. The high crystalline quality of yielded titania originates from in-situ

dissolution–recrystallization in the presence of fluoride ions (reactions (7) and (8)).

45
ACCEPTED MANUSCRIPT

+ ‒ 2‒
4𝐻 + 𝑇𝑖𝑂2 + 6𝐹 → 𝑇𝑖𝐹 6 + 2𝐻2𝑂 (𝑑𝑖𝑠𝑠𝑜𝑙𝑢𝑡𝑖𝑜𝑛) (7)

2‒ + ‒
𝑇𝑖𝐹 6 + 2𝐻2𝑂 →4𝐻 + 𝑇𝑖𝑂2 + 6𝐹 (𝑟𝑒𝑐𝑟𝑦𝑠𝑡𝑎𝑙𝑖𝑧𝑎𝑡𝑖𝑜𝑛) (8)

As noted in Fig. 15a, XRD patterns indicate that the peaks of anatase phase steadily become

stronger and narrower with calcination temperature until T=1200 ˚C, at which transition to

rutile phase is evidenced. The transformation from anatase to rutile normally occurs at much

lower temperatures, around 600 ˚C, which demonstrates the advantage of having F¯ in the

hydrothermal medium on thermal stability. Larger surface area, exposed (001) facets, and

capacity to generate free OH• were believed to be the main contributors to the greater acetone

decomposition rate on T200 with respect to P25 (Fig. 15b).

Fig. 15. (a) XRD patterns of the prepared TiO2 samples; (b) photocatalytic activity in
degradation of acetone; (c and d) TEM images titania after calcination; In Tx, x refers
to the calcination temperature (Lv et al., 2011).
2.4. TiO2 porous and interconnected architectures (3-D)

Hierarchically porous materials with multiscale porosity have drawn attention in

photocatalysis due to their unique features that result in high pollutant removal rate and

enhanced quantum efficiency (Yu et al., 2007d; Chen et al., 2009; Li et al., 2016). To date,

the methods that have been employed for fabrication of hierarchical semiconductor structures

46
ACCEPTED MANUSCRIPT

can be categorized into: templating strategies, template-free strategies, and post-synthetic

treatments (Li et al., 2016). Most of the reports on the synthesis of hierarchically ordered

macro-/mesoporous titania have focused on templating strategies (Deng et al., 2003; Yu et

al., 2007c; Yu et al., 2007d; Li et al., 2016). Nevertheless, a number of researchers

successfully prepared hierarchically porous TiO2 without using any template or additive.

Collins et al. (Collins et al., 2004), for instance, have succeeded in synthesizing ordered

macroporous titania in the absence of template by coupling sol–gel reactions with microphase

separation, transient hydrodynamic gradients, and time-dependent diffusion gradients.

A tri-modally sponge-like macro-/mesoporous titania was synthesized by hydrothermal

treatment of precipitates of TBOT in water in the absence of any additives (Yu et al., 2007c).

The authors suggested that the formation of such structure is due to the cooperative effect of

hydrodynamic flow of water, and restructure and phase transformation of amorphous TiO2

under hydrothermal conditions. The prepared TiO2 samples possess a disordered worm-like

macroporous frameworks with continuous walls made from numerous tightly packed

aggregates of TiO2 nanocrystals. These 3-D continuous macro-porous networks can be

regarded as excellent light transport paths to deliver photons to the interior parts of titania. As

the hydrothermal treatment was extended (from 1 to 36 h), formation of more macropores

and more voluminous mesopores were promoted due to crystal growth and shrinkage of

aggregates. Photocatalytic activity toward acetone consistently enhanced with hydrothermal

duration and reached its maximum at t=24 h and further increment in hydrothermal time to 36

h resulted in reduction in efficiency. The reaction rate constant of the sample treated for 24 h

was about 3 times higher than that of P25, which was attributed to the co-existence of anatase

and brookite, high specific surface area, sufficient crystallinity, and hierarchically porous

structures.

47
ACCEPTED MANUSCRIPT

Employing a similar preparation path, Yu et al. (Yu et al., 2007d) synthesized hierarchical

macro-/mesoporous titania by calcination of precipitates at 300-800 °C instead of

hydrothermal treatment. With increasing the calcination temperature, crystallinity, crystal

size and percentage of rutile phase increased, while surface area, porosity, and photocatalytic

activity plummeted. This was mainly due to the collapse of mesoporous structure and

excessive crystal growth. In particular, after calcination at 600 °C, the sample had a

nonporous structure, implying complete destruction of the mesopores due to the restructuring

of TiO2 during phase transformation from anatase to rutile. A step-by-step formation

procedure was suggested for the hierarchical macro-/mesoporous titania as follows (Fig. 16).

In the absence of stirring, hydrolysis reactions between TBOT droplets and water create a

semipermeable titania membrane at the droplet interface, which compartmentalizes the

following hydrolysis and condensation reactions. Accordingly, via the diffusion of water

through the membrane, hydrolysis and condensation reactions advance inwardly, and almost

perpendicular to the external surface of TBOT droplets. This creates microphase-separated

areas of TiO2 nanoparticles and water/alcohol channels within the TBOT droplets that

undergo spontaneous radial patterning because of the hydrodynamic flow of the solvent.

48
ACCEPTED MANUSCRIPT

Fig. 16. Formation mechanism of the hierarchically porous titania (Yu et al., 2007d).

In another study, Yu et al. (Yu et al., 2007b) explored the effect of hydrothermal temperature

(100-200 ˚C) and time (1-24 h) on the photocatalytic activity and microstructures of bimodal

mesoporous TiO2 powders. Even though the phase transformation temperature from

amorphous to anatase is normally above 400 °C, a high crystallinity for TiO2 samples after

only 3 h of hydrothermal treatment at above 100 °C was observed which may be ascribed to

the non-equilibrium pressure environment in autoclave. The average crystal size and relative

anatase crystallinity considerably increased with applying higher hydrothermal temperatures

and longer times due to Ostwald ripening phenomenon and progression of crystallization

49
ACCEPTED MANUSCRIPT

process (Sekine, 2002). N2 adsorption-desorption analysis clearly indicated the presence of

two length scales in titania structure: i) fine intra-aggregated pore formed between primary

crystallites, and ii) large inter-aggregated pore formed by secondary particles (Fig. 17).

Fig. 17. Formation scheme of bimodal nanocrystalline mesoporous TiO2 powders (Yu et
al., 2007b).

With increasing the hydrothermal time or temperature, photocatalytic activity experienced an

upward trend and reached its optimum values at t=10 h (k= 10.8×10-3 min-1) and T=180 ˚C

(k= 7.0×10-3 min-1) which were remarkably higher than that of P25 (k = 2.85×10-3 min-1).

This could be attributed to the small crystal size, high surface area and bimodal mesoporous

structures of hydrothermally prepared samples, which facilitate the adsorption and desorption

of reactants and products. Further increment in hydrothermal time to 24 h or temperature to

T=200 ˚C led to decline in activity which were justified considering the reduction in BET

surface area and concentration of surface hydroxyl groups (Yu et al., 2007b).

This research group also investigated the impact of pH of hydrothermal reaction medium

(using HCl or NH3.H2O) on the microstructure and activity of mesoporous titania (Yu et al.,

2006a). XRD results revealed that at pH <= 6 anatase and brookite phases co-exist in the

crystalline structure while at higher pH values (pH= 9 and 11) only anatase is present. On the

other hand, owing to the larger quantity of hydroxyl groups in more basic precipitation

medium, hydrolysis of alkoxides was improved and led to the formation of larger crystals and

50
ACCEPTED MANUSCRIPT

increase in crystallinity. TiO2 prepared at pH=9 showed the maximum activity toward

degradation of acetone (2.17 times higher than P25), which could be attributed to the optimal

trade-off between crystallinity and surface area.

As highlighted in section 2.3, the exposed facet of TiO2 can exert considerable influence on

the photocatalytic activity (Liu et al., 2010a). It has been reported that anatase TiO2 with a

high percentage of (001) facets outperformed P25 in OH• generation (Yang et al., 2009), and

the degradation of methylene orange (Yang et al., 2009) and methylene blue (Tian et al.,

2012). In this regard, a two-step hydrothermal synthesis was developed to fabricate

hierarchical flower-like TiO2 with percentage of the (001) facets as high as 87% (Fig. 18)

(Xiang and Yu, 2011). In the first step, titanate nanotubes were prepared by treating P25 in a

basic solution (water and NaOH) under hydrothermal condition at 150 ˚C for 48 h.

Subsequently, nanotubes endured solvothermal reactions in the presence of ethanol and HF in

a Teflon-lined autoclave at 180 ˚C for 24 h and transformed into hierarchical superstructures

self-assembled from TiO2 nanosheets. It is noteworthy that ethanol is of particular importance

as the reaction medium since it acts as a protective capping agent and hinders the growth of

anatase single crystals along the (001) direction. Despite the fact that P25 possesses larger

surface area with respect to the yielded TiO2 (55.1 vs 38.9 m2/g), the latter exhibited higher

acetone removal efficiency after 60 minutes of reaction (25.9% vs 16.4%) which can stem

from the following facts: (i) the inter-meshed nanosheets allow multiple reflections of UV

light which in turn increases the number of photogenerated e--h+ (also confirmed by the UV-

Vis absorption spectra), (ii) the hierarchical porous structure facilitates the transfer of reactant

and products molecules, and (iii) 2-D nanosheets with exposed (001) facets improve the

adsorption of pollutants and reaction rate.

51
ACCEPTED MANUSCRIPT

Fig. 18. SEM images of hierarchical flower-like TiO2. Inset in (a) is a magnified image of
an individual spherical superstructure. Inset in (b) shows a simulated shape of the
anatase TiO2 single crystal (Xiang and Yu, 2011).
Templating strategy has also been applied for preparation of hierarchical nanostructured

materials for air purification. Self-assembly of surfactant under hydrothermal conditions has

been exploited to direct the formation of inorganic phases for synthesizing

macro/mesoporous metal oxides (Deng et al., 2003; Yuan et al., 2003; Wang et al., 2005). For

instance, Wang et al. (Wang et al., 2005) used a nonionic poly(alkylene oxide)-based

surfactant as the structural-directing agent to prepare hierarchically macro/mesoporous titania

(Fig. 19). A short calcination step (4 h at 350-800 °C) was considered after the hydrothermal

treatment in order to remove the surfactant and improve the crystallization. The photocatalyst

calcinated at 350 °C had an intact macro/mesoporous structure and exceeded commercial P25

in ethylene conversion by 20%. UV-vis diffuse reflection spectra revealed that the

photoabsorption capability of titania samples has an inverse correlation with calcination

temperature, which could result from the diminution/destruction of meso- and macropores

structures at higher temperatures. In other words, the interconnected hierarchical bimodal

porous networks can act as light-passage and enable light waves to penetrate inside TiO2

structure and be utilized more efficiently.

52
ACCEPTED MANUSCRIPT

Fig. 19. SEM images of the titanium dioxide monolithic particles calcinated at 350 °C
(Wang et al., 2005).

A few works have been dedicated to hydrothermal fabrication of 3-D nanocomposites of

TiO2 with other metal/non-metal oxides for PCO air purification (Puddu et al., 2007; Chen et

al., 2009; Longo et al., 2014; Cheng et al., 2018; Lee et al., 2018). The purpose behind using

another oxide with titania can be to extend light absorption, to increase the surface area,

and/or to retard charge carriers recombination.

In a recent study, Cu2O semiconductor was coupled with titanium dioxide to achieve a

visible-light responsive photocatalyst for toluene photodegradation (Cheng et al., 2018).

Cu2O small band-gap (2.0 eV) as well as its CB level allows formation of a p-n

heterojunction between TiO2 and Cu2O. The composite was synthesized in two steps:

hydrothermal preparation of TiO2 and TiO2-Cu2O preparation via glucose solvothermal route.

As can be seen in Fig. 20, Cu2O has a cubic shape (approximately 1×1×1 µm3) and a smooth

surface, which, in a way, served as a support for TiO2 nanoparticles. Toluene removal

efficiency after 3 h of visible light irradiation was 63% for TiO2-Cu2O composite and 39%

for P25. This considerable improvement in photocatalytic activity for the composite was

mainly ascribed to the transfer of electrons from Cu2O to TiO2 and holes in the opposite

direction, which lower the possibility of e--h+ recombination.

53
ACCEPTED MANUSCRIPT

Fig. 20. (a) SEM and (b) TEM images of TiO2-Cu2O composite (Cheng et al., 2018).

Longo et al. (Longo et al., 2014) synthesized BiVO4/TiO2 composite by co-precipitation

route followed by hydrothermal/thermal treatment for isopropanol degradation under indoor

illumination. Unexpectedly, in spite of having a larger band gap, hydrothermally-prepared

pure titania considerably outperformed BiVO4/TiO2 composites, which was assigned to the

low surface area, large crystal size, and low photoactivity of BiVO4 phase. WO3 is another

viable candidate to be coupled with TiO2 considering its small band gap, 2.8 eV, and high

surface acidity. The enhancement of gas phase PCO processes with addition of WO3 to TiO2

has been reported for benzene and 2-propanol (Tae Kwon et al., 2000) and toluene (Fuerte et

al., 2002). Puddu et al. (Puddu et al., 2007) proposed a one-step hydrothermal process for

preparation of TiO2/WO3 nanocomposites and evaluated their efficiency for TCE

degradation. The WO3 loading was found to affect the size of the anatase crystals with an

inverse relationship though it did not significantly affect the porosity of the mixed oxides.

The nanocomposites with WO3 contents of 1–5% wt showed up to 42% higher photocatalytic

activity compared to that of TiO2 despite having a lower surface area. This can be mainly

attributed to the improved charge separation when the two semiconductors were coupled.

To alleviate the negative impact of calcination on textural properties, minimize sintering, and

also increase surface area, ZrO2 or SiO2 was utilized as structural stabilizers to fabricate

54
ACCEPTED MANUSCRIPT

hierarchical macro/mesoporous TiO2/SiO2 (TS) and TiO2/ZrO2 (TZ) nanocomposites in the

presence of a non-ionic poly(alkylene oxide)-based surfactant (C16(EO)10, Brij56)) (Chen et

al., 2009). The samples had a honeycomb porous structure with fine particular morphology,

confirming that mesoporousity partly originated from interparticle porosity. On the other

hand, the macroporous channels were usually of one-dimensional orientation, parallel to each

other and orthogonal to one side of the monolithic particles. Introduction of a second phase

into TiO2 structure not only brought about superior thermo–mechanical strength, but also

improved surface acidity due to the presence of strong surface OH groups. Regarding the

thermal stability of TS and TZ composites, XRD results confirmed that even in samples that

calcinated at 800 ˚C, anatase is the dominant phase and crystal size is much smaller with

respect to that of pure titania. TiO2, TS, and TZ photocatalytic activities towards ethylene

followed different trends with calcination temperature. Ethylene removal efficiency over pure

titania sharply declined from 42 to 4% by increasing the calcination temperature from 350 to

800 ˚C, while TS and TZ activities peaked at 500 ˚C (49% removal) and 650 ˚C (43%)

respectively. In another work, Lee et al. (Lee et al., 2018) synthesized TiO2-SiO2 composites

via a one-step hydrothermal method using surfactant-templating, and evaluated their

photocatalytic activity in formaldehyde removal. SEM analysis of TiO2-SiO2 composites

revealed that the morphology is closely connected to the Ti:Si molar ratio. At Ti/Si=1/9, the

composite has an irregular shape with a 2-D channel-structure and plenty pores on its surface,

resembling SBA-15, while at higher ratios, a 3-D spherical porous structure was formed. All

composites possessed large surface areas (419-787.1 m2/g) and mesoporousity, which could

enhance the mass transfer of formaldehyde from the air to the active sites on TiO2-SiO2

photocatalysts. Formaldehyde removal efficiencies over TiO2-SiO2 composite with Ti/Si=1/9

and P25 were 99 and 20%, respectively.

55
ACCEPTED MANUSCRIPT

Summary and outlook

In order to advance PCO acceptability into large-scale application in air purification, it is

crucial to develop more efficient and durable photocatalysts. Hence, synthetic methods with

better control over physical and chemical characteristics of TiO2 are of great importance. In

this regard, hydrothermal/solvothermal technique has shown great promises by overcoming

some of the most critical shortcomings of photocatalysts derived from other methods.

In this article, a comprehensive overview of the experimental works conducted on

hydrothermal preparation of titanium dioxide photocatalysts for air purification has been

provided. TiO2 photocatalysts are categorized based on their structural dimensionality and

investigated from preparation, properties, and performance standpoints. TiO2 nanoparticles

and microspheres (i.e. 0-D structures) benefit from high surface area and light-harvesting

ability, and comparatively more straightforward preparation procedures. Their crystalline and

structural properties can be controlled by pH of the reaction medium, usage of water-alcohol

as solvent, and calcination temperature. Hollow spheres can be successfully fabricated from

various titanium precursors in the presence of F‾ ions under hydrothermal environment.

Coupling nanoparticles with graphene can improve adsorption ability, charge carrier lifetime,

and ultimately pollutant removal efficiency. On the other hand, metal (e.g. Ag and Au) and

non-metal dopants (e.g. N) can shift TiO2 photoresponse into the visible light region and

reduce charge recombination. One-dimensional structures (nanotubes and nanorodes) provide

multiple pathways for diffusion of molecules and low e--h+ recombination rate, resulting in

good photocatalytic activity. 1-D TiO2 photocatalysts can be simply produced from

nanoparticles via hydrothermal treatment in highly alkaline solutions. Two-dimensional

nanosheets with high-energy facets can offer superior adsorption and photocatalytic

performance towards many VOCs with respect to nanoparticles. Hydrofluoric acid as a

capping agent is applied in the majority of works for preparation of TiO2 nanosheets.

56
ACCEPTED MANUSCRIPT

Hierarchical porosity and interconnected structure of 3-D titania facilitate pollutants

adsorption and light diffusion into the interior regions of photocatalyst, enhancing the gas-

phase activity.

In our opinion, several research areas are still lacking in hydrothermal preparation of TiO2

materials and deserve further investigation:

 Systematic studies to elucidate how a particular factor (temperature, time, pH,

solvent, surfactant, additives, etc) affects the properties of final product.

 Developing relationships between photocatalyst properties (e.g. crystallinity, exposed

facet, surface area, hydroxyl group content, etc) and activity.

 Designing humidity-resistant photocatalysts by adjusting the hydrophilicity of TiO2 or

by using hydrophobic support materials.

 Template-free synthesis methods for fabrication of 3-D structures with desired

morphology and multimodal porosity.

Acknowledgement

The authors would like to express their gratitude to Concordia University for the support

through the Concordia Research Chair – Energy & Environment.

References

Ardizzone, S., Bianchi, C.L., Cappelletti, G., Gialanella, S., Pirola, C., Ragaini, V., 2007.
Tailored Anatase/Brookite Nanocrystalline TiO2. The Optimal Particle Features for Liquid-
and Gas-Phase Photocatalytic Reactions. The Journal of Physical Chemistry C 111, 13222-
13231.
Ardizzone, S., Bianchi, C.L., Cappelletti, G., Naldoni, A., Pirola, C., 2008. Photocatalytic
degradation of toluene in the gas phase: Relationship between surface species and catalyst
features. Environmental Science and Technology 42, 6671-6676.
Bian, Z., Zhu, J., Li, H., 2016. Solvothermal alcoholysis synthesis of hierarchical TiO2 with
enhanced activity in environmental and energy photocatalysis. Journal of Photochemistry and
Photobiology C: Photochemistry Reviews 28, 72-86.
Bianchi, C.L., Gatto, S., Pirola, C., Naldoni, A., Di Michele, A., Cerrato, G., Crocellà, V.,
Capucci, V., 2014. Photocatalytic degradation of acetone, acetaldehyde and toluene in gas-

57
ACCEPTED MANUSCRIPT

phase: Comparison between nano and micro-sized TiO2. Applied Catalysis B: Environmental
146, 123-130.
Byrappa, K., Adschiri, T., 2007. Hydrothermal technology for nanotechnology. Progress in
Crystal Growth and Characterization of Materials 53, 117-166.
Cano-Casanova, L., Amorós-Pérez, A., Ouzzine, M., Lillo-Ródenas, M.A., Román-Martínez,
M.C., 2018. One step hydrothermal synthesis of TiO2 with variable HCl concentration:
Detailed characterization and photocatalytic activity in propene oxidation. Applied Catalysis
B: Environmental 220, 645-653.
Caruso, R.A., Susha, A., Caruso, F., 2001. Multilayered Titania, Silica, and Laponite
Nanoparticle Coatings on Polystyrene Colloidal Templates and Resulting Inorganic Hollow
Spheres. Chemistry of Materials 13, 400-409.
Chatterjee, S., Bhattacharyya, K., Ayyub, P., Tyagi, A.K., 2010. Photocatalytic Properties of
One-Dimensional Nanostructured Titanates. The Journal of Physical Chemistry C 114, 9424-
9430.
Chen, K., Zhu, L., Yang, K., 2015a. Acid-assisted hydrothermal synthesis of nanocrystalline
TiO2 from titanate nanotubes: Influence of acids on the photodegradation of gaseous toluene.
Journal of Environmental Sciences 27, 232-240.
Chen, K., Zhu, L., Yang, K., 2015b. Tricrystalline TiO2 with enhanced photocatalytic
activity and durability for removing volatile organic compounds from indoor air. Journal of
Environmental Sciences 32, 189-195.
Chen, Q., Zhou, W., Du, G.H., Peng, L.M., 2002. Trititanate nanotubes made via a single
alkali treatment. Advanced Materials 14, 1208-1211+1178.
Chen, X., Mao, S.S., 2007. Titanium Dioxide Nanomaterials: Synthesis, Properties,
Modifications, and Applications. Chemical Reviews 107, 2891-2959.
Chen, X., Wang, X., Fu, X., 2009. Hierarchical macro/mesoporous TiO2/SiO2 and
TiO2/ZrO2 nanocomposites for environmental photocatalysis. Energy & Environmental
Science 2, 872-877.
Cheng, Y., Gao, X., Zhang, X., Su, J., Wang, G., Wang, L., 2018. Synthesis of a TiO2–Cu2O
composite catalyst with enhanced visible light photocatalytic activity for gas-phase toluene.
New Journal of Chemistry 42, 9252-9259.
Choi, W., 2006. Pure and modified TiO2 photocatalysts and their environmental applications.
Catalysis Surveys from Asia 10, 16-28.
Collins, A., Carriazo, D., Davis, S.A., Mann, S., 2004. Spontaneous template-free assembly
of ordered macroporous titania. Chemical Communications, 568-569.
Demeestere, K., Dewulf, J., Van Langenhove, H., 2007. Heterogeneous Photocatalysis as an
Advanced Oxidation Process for the Abatement of Chlorinated, Monocyclic Aromatic and
Sulfurous Volatile Organic Compounds in Air: State of the Art. Critical Reviews in
Environmental Science and Technology 37, 489-538.
Deng, W., Toepke, M.W., Shanks, B.H., 2003. Surfactant-Assisted Synthesis of Alumina
with Hierarchical Nanopores. Advanced Functional Materials 13, 61-65.
Deng, X., Yue, Y., Gao, Z., 2002. Gas-phase photo-oxidation of organic compounds over
nanosized TiO2 photocatalysts by various preparations. Applied Catalysis B: Environmental
39, 135-147.
Dong, J., Ye, J., Ariyanti, D., Wang, Y., Wei, S., Gao, W., 2018. Enhancing photocatalytic
activities of titanium dioxide via well-dispersed copper nanoparticles. Chemosphere 204,
193-201.
Du, G.H., Chen, Q., Han, P.D., Yu, Y., Peng, L.M., 2003. Potassium titanate nanowires:
Structure, growth, and optical properties. Physical Review B - Condensed Matter and
Materials Physics 67, 353231-353237.

58
ACCEPTED MANUSCRIPT

Du, J., Chen, W., Zhang, C., Liu, Y., Zhao, C., Dai, Y., 2011. Hydrothermal synthesis of
porous TiO2 microspheres and their photocatalytic degradation of gaseous benzene.
Chemical Engineering Journal 170, 53-58.
Ebrahimi, A., Fatemi, S., 2017. Titania-reduced graphene oxide nanocomposite as a
promising visible light-active photocatalyst for continuous degradation of VVOC in air
purification process. Clean Technologies and Environmental Policy 19, 2089-2098.
Einaga, H., Tokura, J., Teraoka, Y., Ito, K., 2015. Kinetic analysis of TiO2-catalyzed
heterogeneous photocatalytic oxidation of ethylene using computational fluid dynamics.
Chemical Engineering Journal 263, 325-335.
Fresno, F., Coronado, J.M., Tudela, D., Soria, J., 2005. Influence of the structural
characteristics of Ti1−xSnxO2 nanoparticles on their photocatalytic activity for the
elimination of methylcyclohexane vapors. Applied Catalysis B: Environmental 55, 159-167.
Fuerte, A., Hernández-Alonso, M.D., Maira, A.J., Martı́nez-Arias, A., Fernández-Garcı́a, M.,
Conesa, J.C., Soria, J., Munuera, G., 2002. Nanosize Ti–W Mixed Oxides: Effect of Doping
Level in the Photocatalytic Degradation of Toluene Using Sunlight-Type Excitation. Journal
of Catalysis 212, 1-9.
Fujishima, A., Rao, T.N., Tryk, D.A., 2000. Titanium dioxide photocatalysis. Journal of
Photochemistry and Photobiology C: Photochemistry Reviews 1, 1-21.
Fujishima, A., Zhang, X., Tryk, D.A., 2008. TiO2 photocatalysis and related surface
phenomena. Surface Science Reports 63, 515-582.
Grabowska, E., Marchelek, M., Klimczuk, T., Trykowski, G., Zaleska-Medynska, A., 2016.
Noble metal modified TiO2 microspheres: Surface properties and photocatalytic activity
under UV–vis and visible light. Journal of Molecular Catalysis A: Chemical 423, 191-206.
Hayashi, H., Hakuta, Y., 2010. Hydrothermal synthesis of metal oxide nanoparticles in
supercritical water. Materials 3, 3794-3817.
Hernández-Alonso, M.D., García-Rodríguez, S., Suárez, S., Portela, R., Sánchez, B.,
Coronado, J.M., 2011. Highly selective one-dimensional TiO2-based nanostructures for air
treatment applications. Applied Catalysis B: Environmental 110, 251-259.
Hurum, D.C., Gray, K.A., Rajh, T., Thurnauer, M.C., 2005. Recombination pathways in the
degussa P25 formulation of TiO 2: Surface versus lattice mechanisms. Journal of Physical
Chemistry B 109, 977-980.
Ismail, A.A., Bahnemann, D.W., 2011. Mesoporous titania photocatalysts: preparation,
characterization and reaction mechanisms. Journal of Materials Chemistry 21, 11686-11707.
Jain, P.K., El-Sayed, M.A., 2007. Surface Plasmon Resonance Sensitivity of Metal
Nanostructures: Physical Basis and Universal Scaling in Metal Nanoshells. The Journal of
Physical Chemistry C 111, 17451-17454.
Jo, W.-K., Kang, H.-J., 2013. (Ratios: 5, 10, 50, 100, and 200) Polyaniline–TiO2 composites
under visible- or UV-light irradiation for decomposition of organic vapors. Materials
Chemistry and Physics 143, 247-255.
Jo, W.-K., Kang, H.-J., Chun, H.-H., 2014a. Degradation of gas-phase organic contaminants
via nitrogen-embedded one-dimensional rod-shaped titania in a plug-flow reactor.
Environmental Technology 35, 2132-2139.
Jo, W.-K., Lee, J., Chun, H.-H., 2014b. Titania Nanotubes Grown on Carbon Fibers for
Photocatalytic Decomposition of Gas-Phase Aromatic Pollutants. Materials 7, 1801.
Kamat, P.V., 2007. Meeting the Clean Energy Demand: Nanostructure Architectures for
Solar Energy Conversion. The Journal of Physical Chemistry C 111, 2834-2860.
Kasuga, T., Hiramatsu, M., Hoson, A., Sekino, T., Niihara, K., 1998. Formation of Titanium
Oxide Nanotube. Langmuir 14, 3160-3163.

59
ACCEPTED MANUSCRIPT

Kibanova, D., Sleiman, M., Cervini-Silva, J., Destaillats, H., 2012. Adsorption and
photocatalytic oxidation of formaldehyde on a clay-TiO2 composite. Journal of Hazardous
Materials 211-212, 233-239.
Kibanova, D., Trejo, M., Destaillats, H., Cervini-Silva, J., 2009. Synthesis of hectorite–TiO2
and kaolinite–TiO2 nanocomposites with photocatalytic activity for the degradation of model
air pollutants. Applied Clay Science 42, 563-568.
Lazzeri, M., Vittadini, A., Selloni, A., 2001. Structure and energetics of stoichiometric
${\mathrm{TiO}}_{2}$ anatase surfaces. Physical Review B 63, 155409.
Lee, S.-S., Lu, C.-Y., Wu, M.-C., 2018. Study of the structure and characteristics of
mesoporous TiO2 photocatalyst, and evaluation of its factors on gaseous formaldehyde
removal by the analysis of ANOVA and S/N ratio. RSC Advances 8, 22199-22215.
Lewandowski, M., Ollis, D.F., 2003. Halide acid pretreatments of photocatalysts for
oxidation of aromatic air contaminants: rate enhancement, rate inhibition, and a
thermodynamic rationale. Journal of Catalysis 217, 38-46.
Li, H., Wu, X., Yin, S., Katsumata, K., Wang, Y., 2017. Effect of rutile TiO2 on the
photocatalytic performance of g-C3N4/brookite-TiO2-xNy photocatalyst for NO
decomposition. Applied Surface Science 392, 531-539.
Li, X., Yu, J., Jaroniec, M., 2016. Hierarchical photocatalysts. Chemical Society Reviews 45,
2603-2636.
Li, X.Z., Li, F.B., 2001. Study of Au/Au3+-TiO2 Photocatalysts toward Visible
Photooxidation for Water and Wastewater Treatment. Environmental Science & Technology
35, 2381-2387.
Liang, L., Li, K., Lv, K., Ho, W., Duan, Y., 2017. Highly photoreactive TiO2 hollow
microspheres with super thermal stability for acetone oxidation. Chinese Journal of Catalysis
38, 2085-2093.
Liu, C., Han, X., Xie, S., Kuang, Q., Wang, X., Jin, M., Xie, Z., Zheng, L., 2013. Enhancing
the Photocatalytic Activity of Anatase TiO2 by Improving the Specific Facet-Induced
Spontaneous Separation of Photogenerated Electrons and Holes. Chemistry – An Asian
Journal 8, 282-289.
Liu, M., Piao, L., Lu, W., Ju, S., Zhao, L., Zhou, C., Li, H., Wang, W., 2010a. Flower-like
TiO2 nanostructures with exposed {001} facets: Facile synthesis and enhanced
photocatalysis. Nanoscale 2, 1115-1117.
Liu, M., Piao, L., Zhao, L., Ju, S., Yan, Z., He, T., Zhou, C., Wang, W., 2010b. Anatase TiO2
single crystals with exposed {001} and {110} facets: facile synthesis and enhanced
photocatalysis. Chemical Communications 46, 1664-1666.
Liu, S., Yu, J., Jaroniec, M., 2011. Anatase TiO2 with Dominant High-Energy {001} Facets:
Synthesis, Properties, and Applications. Chemistry of Materials 23, 4085-4093.
Liu, Z., Zhang, X., Nishimoto, S., Murakami, T., Fujishima, A., 2008. Efficient
Photocatalytic Degradation of Gaseous Acetaldehyde by Highly Ordered TiO2 Nanotube
Arrays. Environmental Science & Technology 42, 8547-8551.
Longo, G., Fresno, F., Gross, S., Štangar, U.L., 2014. Synthesis of BiVO4/TiO2 composites
and evaluation of their photocatalytic activity under indoor illumination. Environmental
Science and Pollution Research 21, 11189-11197.
Lv, K., Xiang, Q., Yu, J., 2011. Effect of calcination temperature on morphology and
photocatalytic activity of anatase TiO2 nanosheets with exposed {001} facets. Applied
Catalysis B: Environmental 104, 275-281.
Lv, Y., Yu, L., Huang, H., Liu, H., Feng, Y., 2009. Preparation of F-doped titania
nanoparticles with a highly thermally stable anatase phase by alcoholysis of TiCl4. Applied
Surface Science 255, 9548-9552.

60
ACCEPTED MANUSCRIPT

Maira, A.J., Coronado, J.M., Augugliaro, V., Yeung, K.L., Conesa, J.C., Soria, J., 2001.
Fourier Transform Infrared Study of the Performance of Nanostructured TiO2 Particles for
the Photocatalytic Oxidation of Gaseous Toluene. Journal of Catalysis 202, 413-420.
Maira, A.J., Yeung, K.L., Lee, C.Y., Yue, P.L., Chan, C.K., 2000. Size Effects in Gas-Phase
Photo-oxidation of Trichloroethylene Using Nanometer-Sized TiO2 Catalysts. Journal of
Catalysis 192, 185-196.
Mamaghani, A.H., Haghighat, F., Lee, C.-S., 2017. Photocatalytic oxidation technology for
indoor environment air purification: The state-of-the-art. Applied Catalysis B: Environmental
203, 247-269.
Mamaghani, A.H., Haghighat, F., Lee, C.-S., 2018a. Gas phase adsorption of volatile organic
compounds onto titanium dioxide photocatalysts. Chemical Engineering Journal 337, 60-73.
Mamaghani, A.H., Haghighat, F., Lee, C.-S., 2018b. Photocatalytic degradation of VOCs on
various commercial titanium dioxides: Impact of operating parameters on removal efficiency
and by-products generation. Building and Environment 138, 275-282.
Mattsson, A., Leideborg, M., Larsson, K., Westin, G., Österlund, L., 2006. Adsorption and
Solar Light Decomposition of Acetone on Anatase TiO2 and Niobium Doped TiO2 Thin
Films. The Journal of Physical Chemistry B 110, 1210-1220.
Minero, C., Mariella, G., Maurino, V., Pelizzetti, E., 2000. Photocatalytic Transformation of
Organic Compounds in the Presence of Inorganic Anions. 1. Hydroxyl-Mediated and Direct
Electron-Transfer Reactions of Phenol on a Titanium Dioxide−Fluoride System. Langmuir
16, 2632-2641.
Mrowetz, M., Selli, E., 2005. Enhanced photocatalytic formation of hydroxyl radicals on
fluorinated TiO2. Physical Chemistry Chemical Physics 7, 1100-1102.
Mrowetz, M., Selli, E., 2006. H2O2 evolution during the photocatalytic degradation of
organic molecules on fluorinated TiO2. New Journal of Chemistry 30, 108-114.
Nakata, K., Fujishima, A., 2012. TiO2 photocatalysis: Design and applications. Journal of
Photochemistry and Photobiology C: Photochemistry Reviews 13, 169-189.
Nguyen, N.H., Bai, H., 2014. Photocatalytic removal of NO and NO2 using titania nanotubes
synthesized by hydrothermal method. Journal of Environmental Sciences 26, 1180-1187.
Pan, J.H., Zhang, X., Du, A.J., Sun, D.D., Leckie, J.O., 2008. Self-Etching Reconstruction of
Hierarchically Mesoporous F-TiO2 Hollow Microspherical Photocatalyst for Concurrent
Membrane Water Purifications. Journal of the American Chemical Society 130, 11256-
11257.
Park, H., Choi, W., 2004. Effects of TiO2 Surface Fluorination on Photocatalytic Reactions
and Photoelectrochemical Behaviors. The Journal of Physical Chemistry B 108, 4086-4093.
Puddu, V., Mokaya, R., Li Puma, G., 2007. Novel one step hydrothermal synthesis of
TiO2/WO3 nanocomposites with enhanced photocatalytic activity. Chemical
Communications, 4749-4751.
Ren, T.-Z., Yuan, Z.-Y., Su, B.-L., 2003. Surfactant-assisted preparation of hollow
microspheres of mesoporous TiO2. Chemical Physics Letters 374, 170-175.
Rendón-Rivera, A., Toledo-Antonio, J.A., Cortés-Jácome, M.A., Angeles-Chávez, C., 2011.
Generation of highly reactive OH groups at the surface of TiO2 nanotubes. Catalysis Today
166, 18-24.
Romero-Vargas Castrillón, S., de Lasa, H.I., 2007. Performance Evaluation of Photocatalytic
Reactors for Air Purification Using Computational Fluid Dynamics (CFD). Industrial &
Engineering Chemistry Research 46, 5867-5880.
Roso, M., Lorenzetti, A., Boaretti, C., Hrelja, D., Modesti, M., 2015. Graphene/TiO2 based
photo-catalysts on nanostructured membranes as a potential active filter media for methanol
gas-phase degradation. Applied Catalysis B: Environmental 176-177, 225-232.

61
ACCEPTED MANUSCRIPT

Roy, N., Sohn, Y., Pradhan, D., 2013. Synergy of Low-Energy {101} and High-Energy
{001} TiO2 Crystal Facets for Enhanced Photocatalysis. ACS Nano 7, 2532-2540.
Sánchez, B., Sánchez-Muñoz, M., Muñoz-Vicente, M., Cobas, G., Portela, R., Suárez, S.,
González, A.E., Rodríguez, N., Amils, R., 2012. Photocatalytic elimination of indoor air
biological and chemical pollution in realistic conditions. Chemosphere 87, 625-630.
Sboui, M., Bouattour, S., Liotta, L.F., Parola, V.L., Gruttadauria, M., Marcì, G., Boufi, S.,
2018. Paper-TiO2 composite: An effective photocatalyst for 2-propanol degradation in gas
phase. Journal of Photochemistry and Photobiology A: Chemistry 350, 142-151.
Sekine, Y., 2002. Oxidative decomposition of formaldehyde by metal oxides at room
temperature. Atmospheric Environment 36, 5543-5547.
Shayegan, Z., Lee, C.-S., Haghighat, F., 2018. TiO2 photocatalyst for removal of volatile
organic compounds in gas phase – A review. Chemical Engineering Journal 334, 2408-2439.
Shengwei, L., Jiaguo, Y., Stephen, M., 2009. Spontaneous construction of photoactive hollow
TiO 2 microspheres and chains. Nanotechnology 20, 325606.
Shi, J., Chen, J., Li, G., An, T., Yamashita, H., 2017. Fabrication of Au/TiO2
nanowires@carbon fiber paper ternary composite for visible-light photocatalytic degradation
of gaseous styrene. Catalysis Today 281, 621-629.
Shi, L., Weng, D., 2008. Highly active mixed-phase TiO2 photocatalysts fabricated at low
temperatureand the correlation between phase compositionand photocatalytic activity.
Journal of Environmental Sciences 20, 1263-1267.
Shibata, T., Sakai, N., Fukuda, K., Ebina, Y., Sasaki, T., 2007. Photocatalytic properties of
titania nanostructured films fabricated from titania nanosheets. Physical Chemistry Chemical
Physics 9, 2413-2420.
Shin, S.-H., Chun, H.-H., Jo, W.-K., 2015. Enhanced Photocatalytic Efficiency of N–F-Co-
Embedded Titania under Visible Light Exposure for Removal of Indoor-Level Pollutants.
Materials 8, 31.
Sofianou, M.-V., Trapalis, C., Psycharis, V., Boukos, N., Vaimakis, T., Yu, J., Wang, W.,
2012. Study of TiO2 anatase nano and microstructures with dominant {001} facets for NO
oxidation. Environmental Science and Pollution Research 19, 3719-3726.
Suave, J., Amorim, S.M., Ângelo, J., Andrade, L., Mendes, A., Moreira, R.F.P.M., 2017.
TiO2/reduced graphene oxide composites for photocatalytic degradation in aqueous and
gaseous medium. Journal of Photochemistry and Photobiology A: Chemistry 348, 326-336.
Subramanian, V., Wolf, E., Kamat, P.V., 2001. Semiconductor−Metal Composite
Nanostructures. To What Extent Do Metal Nanoparticles Improve the Photocatalytic Activity
of TiO2 Films? The Journal of Physical Chemistry B 105, 11439-11446.
Subramanian, V., Wolf, E.E., Kamat, P.V., 2004. Catalysis with TiO2/Gold Nanocomposites.
Effect of Metal Particle Size on the Fermi Level Equilibration. Journal of the American
Chemical Society 126, 4943-4950.
Sun, X., Li, Y., 2003. Synthesis and Characterization of Ion-Exchangeable Titanate
Nanotubes. Chemistry – A European Journal 9, 2229-2238.
Tachikawa, T., Yamashita, S., Majima, T., 2011. Evidence for Crystal-Face-Dependent TiO2
Photocatalysis from Single-Molecule Imaging and Kinetic Analysis. Journal of the American
Chemical Society 133, 7197-7204.
Tae Kwon, Y., Yong Song, K., In Lee, W., Jin Choi, G., Rag Do, Y., 2000. Photocatalytic
Behavior of WO3-Loaded TiO2 in an Oxidation Reaction. Journal of Catalysis 191, 192-199.
Taguchi, T., Saito, Y., Sarukawa, K., Ohno, T., Matsumura, M., 2003. Formation of new
crystal faces on TiO2 particles by treatment with aqueous HF solution or hot sulfuric acid.
New Journal of Chemistry 27, 1304-1306.

62
ACCEPTED MANUSCRIPT

Tian, F., Zhang, Y., Zhang, J., Pan, C., 2012. Raman Spectroscopy: A New Approach to
Measure the Percentage of Anatase TiO2 Exposed (001) Facets. The Journal of Physical
Chemistry C 116, 7515-7519.
Toledo-Antonio, J.A., Capula, S., Cortés-Jácome, M.A., Angeles-Chávez, C., López-Salinas,
E., Ferrat, G., Navarrete, J., Escobar, J., 2007. Low-Temperature FTIR Study of CO
Adsorption on Titania Nanotubes. The Journal of Physical Chemistry C 111, 10799-10805.
Toledo Antonio, J.A., Cortes-Jacome, M.A., Orozco-Cerros, S.L., Montiel-Palacios, E.,
Suarez-Parra, R., Angeles-Chavez, C., Navarete, J., López-Salinas, E., 2010. Assessing
optimal photoactivity on titania nanotubes using different annealing temperatures. Applied
Catalysis B: Environmental 100, 47-54.
van der Meulen, T., Mattson, A., Österlund, L., 2007. A comparative study of the
photocatalytic oxidation of propane on anatase, rutile, and mixed-phase anatase–rutile TiO2
nanoparticles: Role of surface intermediates. Journal of Catalysis 251, 131-144.
Verbruggen, S.W., 2015. TiO2 photocatalysis for the degradation of pollutants in gas phase:
From morphological design to plasmonic enhancement. Journal of Photochemistry and
Photobiology C: Photochemistry Reviews 24, 64-82.
Vijayan, B.K., Dimitrijevic, N.M., Wu, J., Gray, K.A., 2010. The Effects of Pt Doping on the
Structure and Visible Light Photoactivity of Titania Nanotubes. The Journal of Physical
Chemistry C 114, 21262-21269.
Wan, J., Du, X., Wang, R., Liu, E., Jia, J., Bai, X., Hu, X., Fan, J., 2018. Mesoporous
nanoplate multi-directional assembled Bi2WO6 for high efficient photocatalytic oxidation of
NO. Chemosphere 193, 737-744.
Wang, M., Zhang, F., Zhu, X., Qi, Z., Hong, B., Ding, J., Bao, J., Sun, S., Gao, C., 2015.
DRIFTS Evidence for Facet-Dependent Adsorption of Gaseous Toluene on TiO2 with
Relative Photocatalytic Properties. Langmuir 31, 1730-1736.
Wang, R., Cai, X., Shen, F., 2013. Preparation of TiO2 hollow microspheres by a novel
vesicle template method and their enhanced photocatalytic properties. Ceramics International
39, 9465-9470.
Wang, W., Yu, J., Xiang, Q., Cheng, B., 2012. Enhanced photocatalytic activity of
hierarchical macro/mesoporous TiO2–graphene composites for photodegradation of acetone
in air. Applied Catalysis B: Environmental 119-120, 109-116.
Wang, X., Yu, J.C., Ho, C., Hou, Y., Fu, X., 2005. Photocatalytic Activity of a Hierarchically
Macro/Mesoporous Titania. Langmuir 21, 2552-2559.
Wu, C., Yue, Y., Deng, X., Hua, W., Gao, Z., 2004. Investigation on the synergetic effect
between anatase and rutile nanoparticles in gas-phase photocatalytic oxidations. Catalysis
Today 93-95, 863-869.
Wu, H., Ma, J., Li, Y., Zhang, C., He, H., 2014. Photocatalytic oxidation of gaseous ammonia
over fluorinated TiO2 with exposed (001) facets. Applied Catalysis B: Environmental 152-
153, 82-87.
Wu, L., Yu, J.C., Wang, X., Zhang, L., Yu, J., 2005. Characterization of mesoporous
nanocrystalline TiO2 photocatalysts synthesized via a sol-solvothermal process at a low
temperature. Journal of Solid State Chemistry 178, 321-328.
Wu, X.-F., Song, H.-Y., Yoon, J.-M., Yu, Y.-T., Chen, Y.-F., 2009. Synthesis of Core−Shell
Au@TiO2 Nanoparticles with Truncated Wedge-Shaped Morphology and Their
Photocatalytic Properties. Langmuir 25, 6438-6447.
Wu, Z., Gu, Z., Zhao, W., Wang, H., 2007. Photocatalytic oxidation of gaseous benzene over
nanosized TiO2 prepared by solvothermal method. Chinese Science Bulletin 52, 3061-3067.
Xiang, Q., Lv, K., Yu, J., 2010. Pivotal role of fluorine in enhanced photocatalytic activity of
anatase TiO2 nanosheets with dominant (001) facets for the photocatalytic degradation of
acetone in air. Applied Catalysis B: Environmental 96, 557-564.

63
ACCEPTED MANUSCRIPT

Xiang, Q., Yu, J., 2011. Photocatalytic Activity of Hierarchical Flower-Like TiO2
Superstructures with Dominant {001} Facets. Chinese Journal of Catalysis 32, 525-531.
Xiao, J., Xie, Y., Cao, H., 2015. Organic pollutants removal in wastewater by heterogeneous
photocatalytic ozonation. Chemosphere 121, 1-17.
Xu, D., Li, J., Yu, Y., Li, J., 2012. From titanates to TiO2 nanostructures: Controllable
synthesis, growth mechanism, and applications. Science China Chemistry 55, 2334-2345.
Yadav, H.M., Kim, J.-S., 2016. Solvothermal synthesis of anatase TiO2-graphene oxide
nanocomposites and their photocatalytic performance. Journal of Alloys and Compounds
688, 123-129.
Yang, H.G., Liu, G., Qiao, S.Z., Sun, C.H., Jin, Y.G., Smith, S.C., Zou, J., Cheng, H.M., Lu,
G.Q., 2009. Solvothermal Synthesis and Photoreactivity of Anatase TiO2 Nanosheets with
Dominant {001} Facets. Journal of the American Chemical Society 131, 4078-4083.
Yang, H.G., Zeng, H.C., 2004. Preparation of Hollow Anatase TiO2 Nanospheres via
Ostwald Ripening. The Journal of Physical Chemistry B 108, 3492-3495.
Yao, J., Zhang, Y., Wang, Y., Chen, M., Huang, Y., Cao, J., Ho, W., Lee, S.C., 2017.
Enhanced photocatalytic removal of NO over titania/hydroxyapatite (TiO2/HAp) composites
with improved adsorption and charge mobility ability. RSC Advances 7, 24683-24689.
Yin, Y., Rioux, R.M., Erdonmez, C.K., Hughes, S., Somorjai, G.A., Alivisatos, A.P., 2004.
Formation of Hollow Nanocrystals Through the Nanoscale Kirkendall Effect. Science 304,
711-714.
Yu, H., Yu, J., Cheng, B., 2007a. Photocatalytic activity of the calcined H-titanate nanowires
for photocatalytic oxidation of acetone in air. Chemosphere 66, 2050-2057.
Yu, J., Low, J., Xiao, W., Zhou, P., Jaroniec, M., 2014. Enhanced Photocatalytic CO2-
Reduction Activity of Anatase TiO2 by Coexposed {001} and {101} Facets. Journal of the
American Chemical Society 136, 8839-8842.
Yu, J., Shi, L., 2010. One-pot hydrothermal synthesis and enhanced photocatalytic activity of
trifluoroacetic acid modified TiO2 hollow microspheres. Journal of Molecular Catalysis A:
Chemical 326, 8-14.
Yu, J., Su, Y., Cheng, B., Zhou, M., 2006a. Effects of pH on the microstructures and
photocatalytic activity of mesoporous nanocrystalline titania powders prepared via
hydrothermal method. Journal of Molecular Catalysis A: Chemical 258, 104-112.
Yu, J., Wang, G., Cheng, B., Zhou, M., 2007b. Effects of hydrothermal temperature and time
on the photocatalytic activity and microstructures of bimodal mesoporous TiO2 powders.
Applied Catalysis B: Environmental 69, 171-180.
Yu, J., Wang, W., Cheng, B., Su, B.-L., 2009a. Enhancement of Photocatalytic Activity of
Mesporous TiO2 Powders by Hydrothermal Surface Fluorination Treatment. The Journal of
Physical Chemistry C 113, 6743-6750.
Yu, J., Yu, H., Cheng, B., Trapalis, C., 2006b. Effects of calcination temperature on the
microstructures and photocatalytic activity of titanate nanotubes. Journal of Molecular
Catalysis A: Chemical 249, 135-142.
Yu, J., Yu, H., Cheng, B., Zhao, X., Zhang, Q., 2006c. Preparation and photocatalytic activity
of mesoporous anatase TiO2 nanofibers by a hydrothermal method. Journal of
Photochemistry and Photobiology A: Chemistry 182, 121-127.
Yu, J., Yue, L., Liu, S., Huang, B., Zhang, X., 2009b. Hydrothermal preparation and
photocatalytic activity of mesoporous Au–TiO2 nanocomposite microspheres. Journal of
Colloid and Interface Science 334, 58-64.
Yu, J., Zhang, J., 2010. A simple template-free approach to TiO2 hollow spheres with
enhanced photocatalytic activity. Dalton Transactions 39, 5860-5867.

64
ACCEPTED MANUSCRIPT

Yu, J., Zhang, L., Cheng, B., Su, Y., 2007c. Hydrothermal Preparation and Photocatalytic
Activity of Hierarchically Sponge-like Macro-/Mesoporous Titania. The Journal of Physical
Chemistry C 111, 10582-10589.
Yu, J.C., Ho, W., Yu, J., Hark, S.K., Iu, K., 2003. Effects of Trifluoroacetic Acid
Modification on the Surface Microstructures and Photocatalytic Activity of Mesoporous
TiO2 Thin Films. Langmuir 19, 3889-3896.
Yu, J.G., Guo, H., Davis, S.A., Mann, S., 2006d. Fabrication of Hollow Inorganic
Microspheres by Chemically Induced Self-Transformation. Advanced Functional Materials
16, 2035-2041.
Yu, J.G., Su, Y.R., Cheng, B., 2007d. Template-Free Fabrication and Enhanced
Photocatalytic Activity of Hierarchical Macro-/Mesoporous Titania. Advanced Functional
Materials 17, 1984-1990.
Yu, L., Wang, L., Sun, X., Ye, D., 2018. Enhanced photocatalytic activity of rGO/TiO2 for
the decomposition of formaldehyde under visible light irradiation. Journal of Environmental
Sciences 73, 138-146.
Yu, X., Yu, J., Cheng, B., Huang, B., 2009c. One-Pot Template-Free Synthesis of
Monodisperse Zinc Sulfide Hollow Spheres and Their Photocatalytic Properties. Chemistry –
A European Journal 15, 6731-6739.
Yu, Y., Yu, J.C., Yu, J.-G., Kwok, Y.-C., Che, Y.-K., Zhao, J.-C., Ding, L., Ge, W.-K.,
Wong, P.-K., 2005. Enhancement of photocatalytic activity of mesoporous TiO2 by using
carbon nanotubes. Applied Catalysis A: General 289, 186-196.
Yuan, Z.Y., Ren, T.Z., Su, B.L., 2003. Hierarchically Mesostructured Titania Materials with
an Unusual Interior Macroporous Structure. Advanced Materials 15, 1462-1465.
Yuan, Z.Y., Zhang, X.B., Su, B.L., 2004. Moderate hydrothermal synthesis of potassium
titanate nanowires. Applied Physics A: Materials Science and Processing 78, 1063-1066.
Zeng, Q., Xie, X., Wang, X., Wang, Y., Lu, G., Pui, D.Y.H., Sun, J., 2018. Enhanced
photocatalytic performance of Ag@TiO2 for the gaseous acetaldehyde photodegradation
under fluorescent lamp. Chemical Engineering Journal 341, 83-92.
Zhang, K., Zhang, X., Chen, H., Chen, X., Zheng, L., Zhang, J., Yang, B., 2004. Hollow
Titania Spheres with Movable Silica Spheres Inside. Langmuir 20, 11312-11314.
Zhang, L., Tse, M.S., Tan, O.K., 2014. Facile in situ synthesis of visible light-active
Pt/C−TiO2 nanoparticles for environmental remediation. Journal of Environmental Chemical
Engineering 2, 1214-1220.
Zhang, Q., Wang, W., Goebl, J., Yin, Y., 2009. Self-templated synthesis of hollow
nanostructures. Nano Today 4, 494-507.
Zhang, Y., Tang, Z.-R., Fu, X., Xu, Y.-J., 2010. TiO2−Graphene Nanocomposites for Gas-
Phase Photocatalytic Degradation of Volatile Aromatic Pollutant: Is TiO2−Graphene Truly
Different from Other TiO2−Carbon Composite Materials? ACS Nano 4, 7303-7314.
Zhao, T., Liu, Z., Nakata, K., Nishimoto, S., Murakami, T., Zhao, Y., Jiang, L., Fujishima,
A., 2010. Multichannel TiO2 hollow fibers with enhanced photocatalytic activity. Journal of
Materials Chemistry 20, 5095-5099.
Zhao, W., Dai, J., Liu, F., Bao, J., Wang, Y., Yang, Y., Yang, Y., Zhao, D., 2012.
Photocatalytic oxidation of indoor toluene: Process risk analysis and influence of relative
humidity, photocatalysts, and VUV irradiation. Science of The Total Environment 438, 201-
209.
Zheng, Y., Cai, J., Lv, K., Sun, J., Ye, H., Li, M., 2014. Hydrogen peroxide assisted rapid
synthesis of TiO2 hollow microspheres with enhanced photocatalytic activity. Applied
Catalysis B: Environmental 147, 789-795.
Zhong, L., Haghighat, F., 2015. Photocatalytic air cleaners and materials technologies –
Abilities and limitations. Building and Environment 91, 191-203.

65
ACCEPTED MANUSCRIPT

Zhong, Z., Yin, Y., Gates, B., Xia, Y., 2000. Preparation of mesoscale hollow spheres of
TiO2 and SnO2 by templating against crystalline arrays of polystyrene beads. Advanced
Materials 12, 206-209.
Zhou, M., Xu, J., Yu, H., Liu, S., 2010. Low-temperature hydrothermal synthesis of highly
photoactive mesoporous spherical TiO2 nanocrystalline. Journal of Physics and Chemistry of
Solids 71, 507-510.
Zhu, Y., Mei, T., Wang, Y., Qian, Y., 2011. Formation and morphology control of
nanoparticlesvia solution routes in an autoclave. Journal of Materials Chemistry 21, 11457-
11463.

66
ACCEPTED MANUSCRIPT

Highlights

 Highly efficient TiO2 photocatalysts can be synthesized via hydrothermal method.

 Morphology and physico-chemical properties of TiO2 govern the performance of PCO air

cleaners.

 Photocatalyst features can be finely tuned during hydrothermal synthesis to optimize the

activity.

 Modifying TiO2 with graphene oxide or dopants boost photocatalytic reactions efficiency.

 Hierarchically porous TiO2 materials offer superior adsorption and light utilization properties.

You might also like