1 s2.0 S1389556715000271 Main

Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42

Contents lists available at ScienceDirect

Journal of Photochemistry and Photobiology C:


Photochemistry Reviews
journal homepage: www.elsevier.com/locate/jphotochemrev

Review

Review of material design and reactor engineering on TiO2


photocatalysis for CO2 reduction
Oluwafunmilola Ola* , M.Mercedes Maroto-Valer
Centre for Innovation in Carbon Capture and Storage (CICCS), School of Engineering and Physical Sciences, Heriot-Watt University, Edinburgh EH14 4AS,
United Kingdom

A R T I C L E I N F O A B S T R A C T

Article history: The continuous combustion of non-renewable fossil fuels and depletion of existing resources is
Received 25 February 2015 intensifying the research and development of alternative future energy options that can directly abate
Received in revised form 19 May 2015 and process ever-increasing carbon dioxide (CO2) emissions. Since CO2 is a thermodynamically stable
Accepted 10 June 2015
compound, its reduction must not consume additional energy or increase net CO2 emissions. Renewable
Available online 23 June 2015
sources like solar energy provide readily available and continuous light supply required for driving this
conversion process. Therefore, the use of solar energy to drive CO2 photocatalytic reactions
Keywords:
simultaneously addresses the aforementioned challenges, while producing sustainable fuels or
Co2 photoreduction
Photocatalysis
chemicals suitable for use in existing energy infrastructure. Recent progress in this area has focused
Photocatalytic reactor on the development and testing of promising TiO2 based photocatalysts in different reactor
Solar fuel production configurations due to their unique physicochemical properties for CO2 photoreduction. TiO2
Titanium dioxide nanostructured materials with different morphological and textural properties modified by using
Artificial photosynthesis organic and inorganic compounds as photosensitizers (dye sensitization), coupling semiconductors of
different energy levels or doping with metals or non-metals have been tested. This review presents
contemporary views on state of the art in photocatalytic CO2 reduction over titanium oxide (TiO2)
nanostructured materials, with emphasis on material design and reactor configurations. In this review,
we discuss existing and recent TiO2 based supports, encompassing comparative analysis of existing
systems, novel designs being employed to improve selectivity and photoconversion rates as well as
emerging opportunities for future development, crucial to the field of CO2 photocatalytic reduction. The
influence of different operating and morphological variables on the selectivity and efficiency of CO2
photoreduction is reviewed. Finally, perspectives on the progress of TiO2 induced photocatalysis for CO2
photoreduction will be presented.
ã 2015 Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/
licenses/by/4.0/).

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.1. Scope of the review . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.2. CO2 photocatalysis . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.3. TiO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2. Modified TiO2 catalysts . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.1. Dye sensitization . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.2. Coupling of semiconductors . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3. Metal and non-metal modifications . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.3.1. Metal doping . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.3.2. Metal semiconductor modification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.3.3. Non-metal modification . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.3.4. Co-doping . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

* Corresponding author at: School of Engineering and Physical Sciences, Heriot-


Watt University, Edinburgh, United Kingdom.
E-mail address: O.O.Ola@hw.ac.uk (O. Ola).

http://dx.doi.org/10.1016/j.jphotochemrev.2015.06.001
1389-5567/ ã 2015 Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42 17

3. Influence of operating parameters on CO2 reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28


3.1. Effect of reductant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.2. Effect of temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3. Effect of pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.4. Effect of particle size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4. Catalyst configuration: supports . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.1. Glass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.2. Optical fibers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.3. Monoliths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.4. Other supports . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
5. Support immobilization techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
5.1. Sol–gel method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
5.1.1. Thermal treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
5.1.2. Influence of organic contaminants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
5.2. Vapor deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
6. Photoreactor design and configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
6.1. Fluidized and slurry reactor designs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
6.2. Fixed bed reactor designs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
7. Conclusions and future perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

and awards, including 2013 Hong Kong University William Mong Distin-
Dr Oluwafunmilola Ola joined the Centre for guished Lecture, 2011 RSC Environment, Sustainability and Energy Division
Innovation in Carbon Capture and Storage at Early Career Award, 2009 Philip Leverhulme Prize, 2005 U.S. Department of
Heriot-Watt University as Research Associate in Energy Award for Innovative Development, 1997 Ritchie Prize, 1996 Glenn
CO2 conversion and solar fuels. She is currently Award—Fuel Chemistry Division of the American Chemical Society and the
working on an EPSRC funded project to engineer 1993 ICI Chemical & Polymers Group Andersonian Centenary Prize.
novel photoreactors that can achieve efficient
hydrocarbon conversion and separation from CO2
for solar fuel production. Prior to this, she obtained
MSc. (with distinction in Environmental Engineer-
ing at the University of Nottingham in 2010 and Ph.
D. in Chemical Engineering at Heriot-Watt Univer-
sity. Her work on solar fuel production from CO2 has
resulted in over 25 (7 journals and 19 conference
papers) publications that have been cited over
46 times. She is also a reviewer for 3 journals
(Catalysis Science and Technology, Renewable Ener-
gy and ChemCatChem). Her contribution to this research field within the
interface of materials chemistry and chemical engineering has led to the award
of 10 travel grants and prizes such as RSC Energy Sector Ph.D. Thesis Award,
RSC Solar Fuels Symposium Best Poster Prize, UKERC 3rd Place Poster Award
and Engineering Research Showcase Highest Merit for Poster Presentation 1. Introduction
Award.
1.1. Scope of the review

Prof M. Mercedes Maroto-Valer is the first Robert The pressures arising from the need for improved living
Buchan Chair in Sustainable Energy Engineering at
Heriot-Watt University. This is a joint appointment
standards triggered mainly by economic and population growth
between the School of Engineering and Physical have detrimental effects on the environment through the
Sciences and the Institute of Petroleum Engineering. continuous consumption of finite fossil fuel reserves linked to
At Heriot-Watt, she is the Head of the Institute for
Mechanical, Processing and Energy Engineering
increasing CO2 emissions [1,2]. The need to meet global energy
(School of Engineering and Physical Sciences) and demand predicted to increase due to a rising global population has
leads the pan-University Energy Academy. She is led to the development of different strategies by which CO2
also Director of the EPSRC funded Centre for
emissions can be reduced. These can be achieved through the use
Innovation in Carbon Capture and Storage (CICCS).
She is a member of the Directorate of the Scottish of non-fossil fuels such as hydrogen, renewable and nuclear
Carbon Capture and Storage (SCCS).She obtained a energy, increased energy efficiency, reduced deforestation and
BSc with Honours (First Class) in Applied Chemistry capture and storage of CO2 emissions or by using a combination of
in 1993 and then a Ph.D. in 1997 at the University of
Strathclyde (Scotland). Following a one-year post- these options. Although nuclear energy can supply low carbon
doctoral fellowship at the Centre for Applied Energy Research (CAER) at the energy, there are concerns with regards to waste generation [3,4].
University of Kentucky in US, she moved to the Pennsylvania State University Public acceptance and limited water availability are also key issues
in US, where she worked as Research Fellow and from 2001 as Assistant
Professor and became Program Coordinator for Sustainable Energy. She joined associated with this technology. There are several challenges,
the University of Nottingham as Reader in 2005 and within 3 years she was including capital cost, source and seasonal availability, economic
promoted to Professor in Energy Technologies. During her time at Nottingham barriers, geographical distribution and environmental issues as
she was the Head of the Energy and Sustainability Research Division at the
Faculty of Engineering.She has over 300 publications, including editor of
major constraints in the use of renewable energy like biomass,
3 books, 90 refereed publications in journals and book chapters and over hydropower, solar and wind energy [1,5].
200 contributions to other journals and conference proceedings. She holds Conversely, the use of hydrogen energy eliminates the
leading positions in professional societies and editorial boards, including FRSC,
Chair of the RSC Energy Sector, Member of the Scientific Advisory Committee
constraints associated with environmental impacts, but requires
of the RCUK Energy Programme and of the UKERC Research Committee and full optimization and energy input as its production is mainly from
Editor-in-Chief of Greenhouse Gases: Science and Technology.Her outstanding steam reforming and water electrolysis. Additional drawbacks are
contributions, publication record and service to the chemical sciences and
related to storage and hydrogen fueling infrastructures such as fuel
energy engineering have been recognized with numerous international prizes
18 O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42

cell vehicles and fueling stations which are still being developed stability. However, its use is limited due to its large band gap; as it
[6]. Carbon dioxide capture and storage (CCS) serve as a means of can only be activated by ultraviolet (UV) light which represents
reducing CO2 emissions, where CO2 is removed and captured from 2–5% of sunlight [19]. Attempts to improve the efficiency of this
large point sources of industrial processes, such as petrochemical catalyst for CO2 photocatalysis are limited to the overall process
plants, power generation, cement, iron and steel production and efficiency being largely dependent on two factors—the physico-
others. This is followed by its subsequent transport, injection and chemical properties of the catalyst and reactor configuration. The
storage in various sinks, such as geological storage (underground optical and electronic properties of TiO2 can be modified through
saline aquifers, depleted oil and gas reservoirs and deep coal the addition of metals or their oxides such as Cu [20], Ag [21], Pd
seams), mineral carbonation and ocean storage. The CO2 separated [22] and Rh [23] and non-metals, such as nitrogen [24] and iodine
and captured is then considered to be stored for prolonged periods [25]. When these metal and non-metal atoms occupy the
ranging from centuries to millions of years [7]. Although the interstitial sites, replace Ti in the substitutional sites or form
technologies associated with CO2 capture and separation show aggregates on the surface of TiO2, they can cause changes in the
great potential with regards to reducing cost and energy penalty properties of TiO2 [26], where the band structures and properties
[8], they still require further development. Bachu et al. compared of TiO2 have been reported to be tailored by this process. These
the aforementioned storage options and identified geological metals also serve as a source of charge-carrier traps which can
storage as the preferable option owing to the significant quantity of increase the life span of separated electron–hole pairs, and thus
CO2 that can be sequestered, long retention time and great depth of enhancing the efficiency and product selectivity for CO2 photore-
experience from the oil and gas industry that would accelerate the duction [13]. Furthermore, the textural properties such as the
immediate deployment after full-scale implementation [9]. CCS surface and bulk crystal structure, particle size and morphology
still needs optimization in order to fill the gap in knowledge with can also be modified. However, it still remains largely unknown
regards to the location and capacity of possible geological locations how the interaction of metal dopants or their oxides modify the
and possible leakage that could occur during or after injection. surface chemistry and reaction mechanism of TiO2 for CO2
Public acceptance has also been recognized as a key factor that can photoreduction. The configuration of catalyst particles in a
pose barriers to the implementation of geological storage as the photoreactor system is also another factor that can influence the
public can accelerate CCS development [9,10]. overall photocatalytic efficiency of TiO2 [27,28]. This review
As CCS is still in the demonstration phase and may be discusses the current conditions, limitations, correlations and
uneconomical for emissions from small to medium sized sources, possibilities of existing systems i.e., photocatalysts and reactors.
other sustainable alternatives with little or no environmental The concept and mechanism of CO2 photocatalysis using titanium
impacts and zero CO2 emissions need to be developed. These dioxide (TiO2) is presented in Section 1. Section 2 is focused on the
technologies which can offset the cost associated with CO2 capture route by which the physicochemical properties of TiO2 can be
and utilize CO2 for chemicals and fuels production rely solely on modified. The influence of different operational variables i.e.,
technological breakthroughs and market competitiveness due to reductant, temperature, pressure and particle size, on the photo-
their versatile applications. At present, utilization of CO2 accounts activity of TiO2 is addressed in Section 3 while Section 4 reviews
for approximately 2% of emissions and forecasts predict 700 mega- various techniques for the fabrication of immobilized semicon-
tons of CO2/year could be mitigated [11]. Alternatives processes ductor photocatalysts. Section 5 covers different catalyst config-
such as photocatalysis, direct photolysis, and electrochemical urations by which the textural properties of TiO2 can be enhanced.
reduction can utilize CO2 as opposed to geological storage [12,13]. Section 6 describes the current conditions, limitations and
The separated CO2 stream from the capture plant will serve as a possibilities of existing photoreactor configurations available for
feedstock for these conversion methods. The synthesized products CO2 photocatalysis. Comparative analyses of existing systems
such as methane, methanol, ethanol etc., can be used as chemicals, crucial to the field of CO2 photocatalytic reduction are discussed
feedstock in fuel cells or hydrogen sources for electricity. under Sections 4–6. Finally, Section 7 summarizes the review with
Mikkelsen et al. have highlighted the difference between photo- future projections required for driving the field of CO2 photo-
catalysis and electrochemical reduction as the source of electrons catalysis.
which is obtained from irradiating semiconductors under light in
the former and the application of an applied current in the latter 1.2. CO2 photocatalysis
[14]. Although electrochemical cells can convert CO2, Yano et al.
reported low efficiencies resulting from the deactivation of Since CO2 is a chemically stable compound due to its carbon–
electrodes as a major drawback. This is due to the deposition of oxygen bonds (bond enthalpy of C¼O in CO2 is +805 kJ/mol), its
poisoning species, i.e., adsorbed organic compounds, on the conversion to carbon based fuels requires substantial energy input
electrode [15,16]. The need for an inexpensive hydrogen source for bond cleavage [29]. Renewable carbon free sources like solar
and high energy photons have been reported as drawbacks in energy provide readily available and continuous energy supply
direct hydrogenation and photolysis [17,18]. required for driving this conversion process. CO2 photocatalysis
In this review, CO2 utilization by direct catalytic conversion of offers the possibility of utilizing captured CO2 to synthesize
CO2 driven by light energy is described. Although CO2 conversion chemicals and fuels with the aid of semiconductor catalyst(s)
to energy rich and chemically useful products is endothermic, under light irradiation. Apart from solar energy, other readily
renewable carbon free sources like solar energy provide readily accessible light sources can be used. Fig. 1 highlights the typical
available and continuous light supply required for driving this photocatalytic process showing band gap formation in a typical
conversion process under ambient conditions. Thus, carbon based semiconductor photocatalyst when exposed to light radiation. As
fuels and chemicals suitable for end-use infrastructure can be shown in Fig. 1, the band gap is the energy region extending from
produced from the conversion of CO2 and water by semiconductor the bottom of the empty conduction band (CB) to the top of the
photocatalysts capable of simultaneously driving chemical reac- occupied valence band (VB). When an electron excited by light
tions and utilizing solar energy. These value added products can be energy migrates from the fully occupied valence band of the
used directly or supplement feed stocks in hydrocarbon production semiconductor located at an energy level (Ev) to a higher energy
or chemical processes. Amongst semiconductor photocatalysts, (Ec) empty conduction band, electron–hole pairs are created if the
titanium dioxide (TiO2) has been frequently used for UV induced absorbed light energy (hv) is greater than or equal to the band gap
photocatalysis due to its abundance, low cost and chemical (Eg) of the semiconductor [12,30]. Eq. (1) presents the formation of
O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42 19

7 are given in Eqs. (4)–(13) below [33,34].

2Hþ þ 2e ! H2 E0 ¼ 0:41 (4)

1
H2 O ! O2 þ 2Hþ þ 2e E0 ¼ 0:82 (5)
2

CO2 þ e ! CO2  E0 ¼ 1:90 (6)

CO2 þ Hþ þ 2e ! HCO2  E0 ¼ 0:49 (7)


Fig. 1. Schematic of TiO2 photocatalyzed reaction where CB and VB represents the
conduction band, and valence band, respectively.
CO2 þ 2Hþ þ 2e ! CO þ H2 O E0 ¼ 0:53 (8)
 +
electron–hole pairs [31] where e , hv and h represents the
conduction band electron, photon energy and hole in the valence
band, respectively. CO2 þ 4Hþ þ 4e ! HCHO þ H2 O E0 ¼ 0:48 (9)
þ
hne þh
Photocatalyst ! (1)
CO2 + 6H+ + 6e = CH3OH + H2O (10)

e + h+ ! heat (2)
CO2 þ 8Hþ þ 8e ! CH4 2H2 O þ H2 O E0 ¼ 0:24 (11)

Eg = Ec  Ev (3)
2CO2 + 8H+ + 12e = C2H4 + 2H2O (12)
Eq. (2) shows that the charge carriers may also recombine in the
surface or bulk before reacting with adsorbed species, dissipating
energy as heat or light while Eq. (3) shows the band gap energy (Eg)
2 CO2 þ 9Hþ þ 12e ! C2 H5 OH þ 3H2 O E0 ¼ 0:33 (13)
which is equal to the difference between the energy of the
conduction (Ec) and the valence band (Ev). The reduction potential
of photo-generated electrons is the energy level at the bottom of The band gaps of some of the most commonly used photo-
conduction band while the energy level at the top of valence band catalysts are shown in Fig. 2 [35–38]. Although some of these
determines the oxidizing ability of photo-generated holes which semiconductor photocatalysts such as hematite (Fe2O3) are low
determine the ability of the semiconductors to undergo oxidations cost and possess suitable band gap energies for visible light
and reductions [19]. The redox potential levels of the adsorbate absorption, they suffer from different limitations. Metal chalco-
species and the band gap energy determine the likelihood and rate genides semiconductors e.g., CdS, PbS, CdSe etc., have been
of the charge transfer processes for electrons and holes [32]. For an reported as being susceptible to photocorrosion and low stability
electron to be donated to the vacant hole, the redox potential level especially in aqueous media [31,39]. The addition of sulphide or
of the donor is thermodynamically required to be above the VB sulfite to the contacting solution has been described to suppress
position of the semiconductor, while that of the acceptor should be photocorrosion. These semiconductors have also been reported to
below the CB position. The reduction potentials for CO2 photore- show some toxicity [40]. Since semiconductors like WO3, Fe2O3
duction with H2O to various products with reference to NHE at pH and SnO2 possess conduction band edge values below the

Fig. 2. Band gap of some photocatalysts with respect to the redox potential of different chemical species measured at pH of 7. Adapted from Refs. [35–38].
20 O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42

hydrogen potential and Gupta and Tripathi pointed out the use of The anatase form is the most suitable for photocatalytic reactions
an external electrical bias as a requirement needed to achieve due to its larger surface area, stability and higher activity compared
hydrogen evolution during water splitting [41]. Fe2O3 has also been to the rutile form [60–63]. The brookite form is not commonly
reported by Fox and Dulay to show lower photoactivity compared accessible, difficult to synthesize and has not been proved for
to TiO2 and ZnO due to corrosion or the formation of short lived photocatalytic reactions [64], [65]. Bouras et al. reported that
ligand-to-metal or metal-to-ligand charge transfer states [42]. The optimal photocatalytic efficiency can be obtained from a mixture of
formation of Zn(OH)2 on the surface of the semiconductor ZnO anatase with a small percentage of rutile through a synergistic
observed from its dissolution in water has been reported by effect between the two crystalline phases as electron hole
Bahnemann et al. to cause instability and deactivation over time recombination is prevented by the creation of energy wells which
[43]. On the other hand, TiO2 appears to be corrosion resistant and serve as an electron trap formed from the lower band gap of rutile
chemically stable [19]. [62]. Although TiO2 has several unique features, its use is limited
Jeyalakshmi et al. also stated that large band gap semi- due to its large band gap (see Table 1); as it can only be activated by
conductors like TiO2 are more suitable for CO2 photoreduction due ultraviolet light which represents 2–5% of sunlight [54,66]. Since
to sufficient positive and negative redox potentials in the VB and visible light accounts for 45% of the solar spectrum [51,67], there is
CB, respectively, compared to smaller band gap semiconductors a need to develop titania based photocatalysts which are active
like CdS where the energy levels of either the VB or CB tend to be under the visible light spectrum.
unsuitable for water oxidation and/or CO2 photoreduction [44].
The photogenerated electrons and holes may recombine to 2. Modified TiO2 catalysts
generate heat energy or become trapped in surface sites, where
reactions with electron accepting or donating species adsorbed on Since the time scale of electron–hole recombination of TiO2 has
the surface of the semiconductor photocatalyst can occur [30,45]. been reported to be higher than the desirable redox reactions [33],
For redox reactions to take place, electron–hole recombination it is crucial to modify the physicochemical properties of TiO2 to
must be minimized. improve process efficiency. Suitable modification of the optical and
Several researchers have studied CO2 photoreduction using electronic properties of TiO2 results in not only the reduction of the
different semiconductors, including single catalysts like TiO2 [46] band width via the incorporation of addition energy levels but
and ZrO2 [47], double catalysts like Cu–Fe/TiO2–SiO2 [48] and increased lifetime of the photogenerated electrons and holes via
Cu–ZnO/Pt–K2Ti6O13 [49], metal and compound oxides such as effective charge carrier separation and suppression of electron–
CuO [50], LaCoO3 [51], Ga2O3 [52] and ATaO3 where A represents hole recombination. Furthermore, the textural properties such as
Na, Li and K [53]. The basic characteristics of an ideal the surface and bulk crystal structure, particle size and morpholo-
semiconductor photocatalyst were reviewed by Refs. [13,54–59]. gy can also be modified. The photocatalytic activity of TiO2 for
These properties include the presence of a large surface area, cost- visible light can be increased by using organic and inorganic
effectiveness, accessibility, resistance to photocorrosion or pro- compounds as photosensitizers (dye sensitization), coupling
duction of toxic by-products and the ability of the redox potentials semiconductors of different energy levels or doping with metals
of the photogenerated valence band and conduction band to be or non-metals to suppress recombination rate and thus increasing
positive and negative in order for the electrons to act as an acceptor quantum yield [54,56,68,69]. Table 2 highlights a summary of
and donor, respectively. The most frequently used semiconductors literature on CO2 photoreduction using TiO2 modifications
are ZnO, CdS, TiO2, WO3 and NiO. However, compared to these [20,22,25,70–121]. All these strategies are described below.
semiconductors, TiO2 still remains the most researched semicon-
ductor photocatalyst due to its availability, chemical stability, low 2.1. Dye sensitization
cost, high photocatalytic activity and resistance to corrosion.
Dye sensitization is a means of increasing absorption toward
1.3. TiO2 the visible light region through the inducement of the photo-
excited dye molecule [19,56,122,123]. Various dyes which harvest
The most common crystalline phases of TiO2 are rutile, anatase visible light that have been used as sensitizers include rhodamine
and brookite. The bulk properties of the crystalline forms of TiO2 B, porphyrins, thionine, rose bengal, erythrosine B etc. [124–126].
are presented in Table 1. Anatase and rutile have lattice parameters Electrons are transferred from the dye molecule to the conduction
(a and c) of 0.3733/0.4584 nm and 0.9370/0.2953 nm, respectively, band of the semiconductor when the energy level of the dye
in the unit cell based on the body centered tetragonal structure. molecule was more negative than the semiconductor. Fig. 3 and

Table 1
Structural properties of crystalline structures of TiO2.

Properties Crystalline forms

Anatase Rutile Brookite


Crystalline structure Tetragonal Tetragonal Rhombohedral
Lattice constants (nm) a = b = 0.3733 a = b = 0.4584 a = 0.5436
c = 0.9370 c = 0.2953 b = 0.9166
c = 0.5135
Bravais lattice Simple, Body centred Simple, body centred Simple
Density (g/cm3) 3.83 4.24 4.17
Melting point ( C) Turning into rutile 1870 Turning into rutile
Boiling point ( C) 2927a – –
Band gap (eV) 3.2 3.0 –
Refractive index (ng) 2.5688 2.9467 2.8090
Standard heat capacity, Cop 55.52 55.60 –
Dielectric constant 55 110–117 78
a
Pressure at pO2 is 101.325 KPa.
O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42 21

Table 2
Modifications of TiO2 for CO2 photoreduction.

Modifications Photocatalyst Light source Reductant Products References


Dye sensitization Dye sensitized (perylene diimide derivatives) Pt 75 W daylight lamp H2O 0.74 mmol/gcatal CH4 [70]
impregnated on TiO2
N3 dye (RuII(2,20 -bipyridyl-4,40 -dicarboxylate)2–Cu Solar concentrator H2O vapor 0.617 mmol/gcatal h CH4 [71]
(0.5 wt%)–Fe (0.5 wt%)/TiO2
1
Ru/RuOx sensitized TiO2 Solar simulator H2O 900 mL h CH4 [72]
N719/TiO2 300 W Xe lamp H2O/2 M NaOH 0.1781 mmol/cm2 CH3OH [73]
0.1292 mmol/cm2 CH2O

Semiconductor CdSe quantum dot (QD)-Pt/TiO2 films 300 W Xe arc lamp, H2O 48 ppmg1 h1 CH4, [74]
coupling 100 mW/cm2 3.3 ppmg1 h1 CH3OH
23.2 wt% AgBr–TiO2 150 W Xe lamp 0.2 M KHCO3 128.56 mmol g CH4 [75]
77.87 mmol g CH3OH
13.28 mmol g C2H5OH
32.14 mmol g CO
PdS quantum dot (QD)-Cu/TiO2 300 W Xe lamp H2O 0.82 mmol g1 h1 CO [76]
0.58 mmol g1 h1 CH4
0.31 mmol g1 h1 C2H6
CdS/TiO2 nanotubes 500 W Xe lamp 0.8 g NaOH 159.5 mmol g/catal CH3OH [77]
Bi2S3/TiO2 nanotubes 2.52 g Na2SO3 224.6 mmol g/catal CH3OH
CeO2/TiO2 SBA-15 300 W Xe lamp H2O <12mmol/g catal CH4 [78]
45 wt% CdS/TiO2 125 W Hg lamp H2O 16 mmol g/catal CH3OH [79]
3 mmol g/catal CH3OH
TiO2/ZrO2 8 W Hg lamp 0.2 mpl/L NaOH 16 mmol g /catal CH4 [80]
175 mmol g /catal H2
GaP/TiO2 1500 W Xe lamp H2O 118 mM/gcatal CH4 [81]

1
Metal doping 0.5 wt% Ru–TiO2 100 W Hg lamp 1 M 2-propanol 200 mmolg-Ti CH4 [82]
250 mmol g-Ti1 H2
TiO2 pellets UVC lamp H2O vapor 0.16 mmol/h H2 [83]
0.25 mmol/h CH4
1 wt% Ag–TiO2 Solar concentrator H2O vapor 4.12 mmol/gcatal h CH3OH [84,85]
3 wt% CuO–TiO2 6 (10 W) UV lamps, H2O 2655 mmol/gcatal CH3OH [20,86]
2450 mW/cm2
2 wt% Cu–TiO2 8 W Hg lamp 0.2 M NaOH 1000 mmol/gcatal CH3OH [87,88]
20 mmol O2
TiO2 pellets 3 UVC lamps H2O 0.25 mmol h1 CH4 [89]
5.2 wt% Ag–TiO2 300 W Hg lamp 0.2 M NaOH >10 mmol/gcatal CH4 + CH3OH [90]
0.15% Pt–TiO2 nanotubes 8 W Hg lamp H2O 4.8 mmol h1/g Ti1 CH4 [91]
Pd/RuO2/TiO2 450 W Xe short arc lamp 0.05 M NaOH 72 ppm HCOO [92]
0.05 M Na2SO3
2 wt% Pd–TiO2 500 W Hg lamp H2O 24.7  108 mol CH4 [22]
2 wt% Cu–TiO2-SBA 15 400 W halide lamp 0.1 M NaOH and 627 mmol g1 h1 CH4 [93]
H2O
0.5 wt% Cu/TiO2–SiO2 Xe arc lamp, 2.4 mW/cm2 H2O 60 mmol g1 h1 CH4, [94]
10 mmol g1 h1 CO
Kaolinite/TiO2 8 W Hg lamp 0.2 M NaOH 4.5 mmol gcatal CH4, [95]
2.5 mmol gcatal CO,
5 mmol gcatal H2
0.1 wt% Y–TiO2 300 W Hg lamp 0.2 M NaOH 384.62 mmol gcatal HCHO [96]
1 wt% and 3 wt% Ce–TiO2 SBA 15 450 W Xe lamp H2O 1 mmol g1 CO [97]
3 wt% Ag–TiO2 8 W Hg lamp H2O 100 mmol/gcatal H2 [98]
6 mmol/gcatal CH4
14 mmol/gcatal C3H6
Ni–TiO2 (0.1 mol%) 6 (3W/cm2) UV lamps H2O 14 mmolgcatal CH4 [99]
La2O3/TiO2 300 W Xe Lamp H2O 4.57 mmol CH4 [100]
CeF3–TiO2 500 W Xe lamp H2O 162 mmolgcatal CH3OH [101]
1.5 wt% NiO–TiO2 200 W Hg lamp H2O 19.51 mmol/gcatal h CH3OH [102]
8.7 at% Pt/9.6 at% Cu–TiO2 AM 1.5G solar simulator H2O >180 ppm/cm2 h H2 [103]
49 ppm/cm2 h CH4
<25ppm/cm2 h CO
Ce–TiO2 8 W Hg lamp 0.2 N NaOH 16 mmol/gcatal CH4 [104]
750 mmol/gcatal H2
Pt/TiO2 500 W Xe Lamp H2O 389.2 ppm H2 [105]
277.2 ppm CH4
12.4 ppm C2H6
785.3 ppm O2
Ti-KIT-6/Si-Ti = 100 300 W UV lamp H2O 4.14 mmol/gcatal h CH4 [106]
2.55 mmol/gcatal h H2
1.45 mmol/gcatal h CO
Pt/SrTiO3–Rh/Pt/CuAlGaO4 AM 1.5G 2 mM FeCl2/FeCl3 0.52 mmol CH3OH [107]
WO3 0.12 mmol H2
5 mmol O2
Pt/SrTiO3–Rh/Pt/CuAlGaO4 AM 1.5G 2 mM FeCl2/FeCl3 8 mmol/g CH3OH [108]
WO3 1 mmol/g H2
12 mmol/g O2
Degussa P25 TiO2 1000 W Xe lamp H2O [109]
22 O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42

Table 2 (Continued)
Modifications Photocatalyst Light source Reductant Products References
0.1 mmol/h CH4
1.4 mmol/h H2
2.7 mmol/h CO
Ag/BaLa4Ti4O15 400 W Hg lamp H 2O 0.7 mmol HCOOH [110]
10 mmol H2
22 mmol CO
16 mmol O2
CeO2/TiO2 500 W Xe lamp H 2O 2.75 mmol/g h H2/CH4 [111]
1.28 mmol/g h O2
Pt/MgO/TiO2 nanotubes 300 W Hg lamp 0.1 mol/L KHCO3 100.22 ppm/h cm2 CH4 [112]
10.4 ppm/h cm2 CO
In/TiO2 500 W Hg lamp H 2O 244 mmolg1 h1 CH4 [113]
81 mmolg1 h1 CO
TiO2 (20%)/KIT6 300 W lamp H 2O 44.56 mmol/g H2 [114]
44.56 mmol/g CH4
1.09 mmol/g CH3OH
120.54 mmol/g CO

Non-metal N doped TiO2/Ni 15 W UV lamp/ 0.2 mol/L NaOH 482 mmolgcatal CH3OH [115]
doping incandescent lamp and Na2SO3
N doped TiO2/Pt–Cu AM 1.5 outdoor sunlight, H 2O 111 ppm/cm2 h (CO,H2, etc.) [116]
75–102 mW/cm2
N doped TiO2 nanotubes 500 W tungsten/ halogen 0.1 N NaOH 1132.6 mmolgcatal CH3OH [117]
lamp 921.6 mmolgcatal HCHO
12475.8 mmolgcatal HCOOH
C doped TiO2 175 W Hg lamp Na2SO3 2610.98 mmolgcatal HCOOH [118]
I doped TiO2 450 W Xe lamp H 2O 2.4 mmolg1 h1 CO [25]
TiO2/N-100 – H 2O 23 mmolg1 h1 CH3OH [119]
g-C3N4–N–TiO2 (CT–70) 300 W Xe Lamp H 2O 14.73 mmol CO [120]
N–TiO2/spirulina 13 W lamp H 2O 144.99 mmol/g H2 [121]
0.48 mmol/g CH4
0.12 mmol/g C2H4
0.17 mmol/g C2H6

photosensitizer must also have high absorption spectrum in the


visible light and infrared regions, with the excitation state having a
long lifetime as well [69,127,128]. The rate of electron injection and
back electron transfer reactions from the dye molecule to the
photocatalyst depend on the characteristics of the dye molecule
and the properties of TiO2 and its interactions with the dye. Gupta
and Tripathi [41] reported TiO2 as an ideal semiconductor for dye
sensitized solar cells due to its stability, high refractive index which
facilitates increased light absorption, high dielectric constant for
electrostatic shielding of the injected electron from the dye
molecule to the electrolyte and suitable conduction band edge
below the energy level of several dyes. Grätzel et al. conducted CO2
photoreduction using Ru/RuOx sensitized TiO2 and obtained
approximately 900 mL h1 of methane using a solar simulator
with light intensity of 0.08 W cm2 [72]. Ozcan et al. also
Fig. 3. Excitation steps with a photosensitizer, where A and D represent the demonstrated the effect of dye sensitized (perylene diimide
electron acceptor and electron donor, respectively. derivatives) Pt impregnated on TiO2 films on CO2 photoreduction
[70]. It was observed that methane production rate was enhanced
Eqs. (14)–(16) illustrate the reactions involved, including photo- to a maximum value of 0.74 mmol/gcatal by adsorbing dye
excitation, injection of electrons and regeneration of dyes, molecules to Pt–TiO2. When Pt was not loaded on TiO2 films,
respectively [122]. inactivity was observed in the presence of dye sensitizers. The
dye !hndye (14) N3 dye (RuII(2,20 -bipyridyl-4,40 -dicarboxylate)2–(NCS)2) coated
with Cu–Fe/TiO2 utilized by Nguyen et al. was found to be capable
of visible light absorption, producing 0.617 mmol/gcatal h of CH4
þ after 5.5 h [71]. Data comparison between this sample and one
dye !TiO2 dye þe (15)
without dye showed the stability of N3 dye sample over a wide
light spectrum. Yuan et al. investigated the photoreduction of CO2
dyeþ þ e ! dye (16) with H2O using a Cu(I) dye sensitized TiO2 based system [129]. The
introduction of the Cu(I) bipyridine complex was reported to be
beneficial for charge separation in TiO2 under full sun illumination
As shown in Fig. 3, the transferred electron reduces the organic (AM 1.5G). Maximum CH4 production rate of ca. 7 mmol/g1 was
electron acceptor (EA) adsorbed on the surface. An ideal observed following 24 h of visible light irradiation with no CH4
photosensitizer must undergo slow backward reactions and fast detected when the pure Cu(I) dye complex was used. However,
electron injection to attain high efficiency [122]. The instability, light and thermal degradation of dye molecules and
O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42 23

disposal of undesired intermediates formed during reactions have and increased charge separation. Sigmund et al. [134] reported that
been reported as a major drawbacks associated with dye the injection of electrons or holes and their direction is dependent
sensitization [41,74,130]. on the Fermi level and the band gap combination of the
semiconductors. The requirements for successful coupling of
2.2. Coupling of semiconductors semiconductors are efficient and fast electron injection; ability
of the small band gap semiconductor to be excited by visible light
During heterojunction formation in semiconductor based with its conduction band being more negative than that of the
photocatalysts, the direction of transfer of photogenerated charge other semiconductor; proper positioning of the Fermi energy level
carriers from the coupled semiconductors will depend on the and insusceptibility of the semiconductors to photocorrosion
position of the CB and VB. TiO2 can be coupled with semi- [56,134].
conductors via direct or indirect Z scheme. In direct Z-scheme, Wang et al. [74] conducted studies of CO2 photocatalytic
spatial charge separation occurs when electrons and holes are reduction over CdSe quantum dot (QD) loaded with Pt impregnat-
injected to CB and VB of different semiconductors in opposite ed TiO2 films. Typical product yields of 48 ppm g1 h1 (methane)
directions (Fig. 4A), while charge separation does not occur in and 3.3 ppm g1 h1 (methanol) were observed in gas phase using
indirect Z-scheme due to electron and hole transfer occurring in visible light irradiation of 420 nm. They observed that charge
the same direction for different semiconductors (Fig. 4B) injection into TiO2 was facilitated by the shift of the conduction
[122,131–133]. The coupling of these semiconductors result in band of CdSe into higher energy which initiated CO2 reduction.
the balance of their Fermi levels (i.e., energy midway between the Process efficiency was also increased via charge separation due to
conduction and the valence band edges) such that electron flow is electron transfer from CdSe to TiO2. No activity was recorded when
from the semiconductor with the higher Fermi level to the one both semiconductors were used independently and using the same
with the lower Fermi level [134]. Excess negative charges are wavelength of light. Recently, ordered mesoporous silica SBA-15/
created in the semiconductor with the lower Fermi level while TiO2 composites with varying ratios of CeO2 were synthesized by
excess positive charges are created in the semiconductor with the Wang et al. [78] for the reduction of CO2 with H2O under simulated
highest Fermi level due to charge transfer. Thus, coupled semi- solar irradiation. The addition of CeO2 was found to not only
conductors benefit from extended band widths in the visible light influence the light harvesting properties of TiO2 toward the visible

Fig. 4. Coupling of TiO2 with semiconductors (SC) illustrating direct (A) and indirect (B) Z-scheme.
24 O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42

light region but enhance the photocatalytic performance as well.


The improved performance was ascribed to the separation of
photogenerated charge carriers induced from the drift of TiO2
electrons to CeO2.
Li et al. [77] utilized either CdS or Bi2S3 in the modification of
the properties of TiO2 nanotubes for CO2 reduction under visible
light irradiation. The addition of either semiconductor was found
to enhance visible light absorbance and photocatalytic activity of
TiO2 nanotubes, with Bi2S3 exhibiting superior activity due to
better surface area and CO2 adsorption. Optimum methanol yields
of 224.6 mmol/gcatal and 159.5 mmol/gcatal was observed using TiO2
nanotubes coated with Bi2S3 and CdS, respectively. Heterojunction
formation between the semiconductors was reported to play a
crucial role in prolonging the lifetime of charge carriers and
preventing electron/hole recombination. Despite its promising Fig. 5. Two dimensional representation of a single TiO2 crystal lattice showing
results, this technique is not widely applied due to its drawbacks. substitutional and interstitial doping.
These include photo-corrosion in aqueous phase which affects the
durability and stability of the catalyst through leaching out of dopants positioning itself in an interstitial site. The ionic radius
dopant [135,136] and difficulty in finding appropriate semicon- ratio of the cation/anion (r+/r) determines the preference of
ductor pairs such that recombination of charge carriers can be cations to occupy certain interstitial sites [140]. Interstitial sites
reduced [137]. may consist of cations with coordination numbers such as 4
(tetrahedral), 6 (octahedral) etc., based on the radius ratio of these
2.3. Metal and non-metal modifications ions. As the values of the ionic radius ratio increase, the number of
anions packed around the cation increase. In tetrahedral holes, the
The optical and electrochemical properties of TiO2 can be cations are packed between planes of anions in close-packed
enhanced by the addition of metal and non-metal ion(s). The band structures if the ionic radius ratio falls within 0.225 and 0.414.
gap and properties of TiO2 have been reported to be modified when Whilst the radius ratio falls within 0.414 and 0.732, if they are
metals or non-metals occupy non-lattice sites (i.e. interstitial), packed in octahedral holes [138].
replace Ti in the substitutional sites or form aggregates on the According to Pagot and Clerjaud [141] and Seebauer and Kratzer
surface of TiO2 [26]. The redox potential of photogenerated charge [142], the local distortion of the crystal lattice can occur in
carriers and visible light absorption will be determined by the substitutional and interstitial doping due to the difference in the
spectral distribution of the modified photocatalysts, which is atomic radii of the dopants compared with the host atoms and
invariably determined by their chemical states [137]. Apart from their chemical affinity with their surrounding atoms. In these
these metals possessing their own catalytic activity, they also serve lattice defects, the change in electric properties is caused by the
as a source of charge-carrier traps which can increase the life span disruption of the chemical bonding between the atoms and
of separated electron–hole pairs, and thus enhance the efficiency distortion of the geometric arrangement of atoms [138]. Vacancies
and product selectivity for CO2 photoreduction [13,32]. may be created during the catalyst preparation process due to
Whether the metal ions are present in the lattice or TiO2 surface impurity atoms hopping from one vacancy to the other, thus
is dependent on two key factors: the preparation procedure where remaining permanently in the substitutional lattice sites after
the amount and homogeneity of the metal ions in its host oxide are calcination. Dopants can be introduced into sol–gel derived
key parameters and the firing temperatures to which the samples samples at molecular level through the mixing of titanium
have been subjected. Diffusion of the metal ions in TiO2 lattice is precursors with soluble dopant compounds. The introduction of
influenced by temperature; with higher temperatures favouring dopants has been found to alter the degree of crystallinity and
diffusion due to high thermal energy of the atoms. The mecha- phase transformation, thereby, subsequently altering the peak
nisms for metal and non-metal modification are described below. heights, areas and relative intensities [143]. Phase transformation
can be facilitated or inhibited by substitutional dopants when
2.3.1. Metal doping cations enter the anatase lattice and cause an increase or decrease
Neamen [138] described doping as the process of adding foreign in the level of oxygen vacancies through valence or reduction/
or impurity atoms into the crystal lattice of a semiconductor oxidation effects. This leads to the subsequent rearrangement of
material. Alterations to the properties of the semiconductor can atoms in the lattice of TiO2 through the substitution of Ti4+ with
occur when controlled amounts of dopants are added to the cations. Conversely, the formation of Ti interstitials may distort the
semiconductor. Fig. 5 shows the schematic representation of these anatase lattice thus restricting the lattice contraction involved in
lattice defects. When these impurity atoms are located at normal the phase transformation to rutile [143]. The reactions of metal
lattice sites i.e., substitution of the host atom occurs, they are doping are described by the following equations, where Mn
referred to as substitutional doping. Substitutional doping can represents the metal ion dopant [56].
occur when one or more of the following criteria are met: the
Mn+ + hn ! M(n+1) + eCB (17)
differences in atomic radii of the atom types are less than 15%,
dopant and host metals have similar crystal structures and
electronegativity or comparable valences to ensure solubility
[139,140]. Conversely, when these impurity atoms are present Mn+ + hn ! M(n1) + h+VB (18)
between normal lattice sites, they are referred to as interstitial
doping i.e., the host atom dislodged from its normal lattice sites is
forced into voids between atoms [138,139]. The likelihood of an Mn+ + eCB M(n1)+ as electron trap (19)
atom occupying an interstitial site can be predicted by comparison
of the radii of the interstitial dopant to the host metal [140]. The
greater difference between these atomic radii results in the Mn+ + h+VB M(n+1)+ as hole trap (20)
O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42 25

Eqs. (17) and (18) depict the formation of energy levels in the band They also reported that charge separation could be improved by co-
gap of TiO2, while (19) and (20) represent the transfer of electrons doping to produce an overall beneficial effect.
between TiO2 and metal ions. However, the energy level (Mn Extensive studies on the improvement of the electronic
+
/M(n1)+) must be less negative than conduction band (CB) of structure of TiO2 have been performed by doping with high
TiO2, while the energy level (Mn+/M(n+1)+) must be less positive energy transition metals by several researchers [130,144–147].
than valence band of TiO2. The influence of iodine doping on the Their results established that the amount and type of metal dopant
phase transformation and photocatalytic activity of TiO2 for CO2 as well as the method of synthesis were key factors in determining
reduction was evaluated by Ref. [25]. The anatase fraction of their photocatalytic activity and the extent of red shift that can be
iodine doped samples was found to increase with increased iodine achieved in the visible light region. Comparisons between the
concentration and calcination temperature while the brookite absorption spectra of metal doped TiO2 synthesized by chemical
fraction was found to decrease under the aforementioned doping (impregnation) and metal ion implantation were made.
conditions. They attributed the optimal visible light activity of Samples synthesized by the latter method exhibited shifts in their
their 5 wt% I–TiO2, calcined at 648 K sample to combinational absorption band toward visible light region (600 nm) caused by
effect of increased surface area, improved visible light absorption intense distance interaction between the metal ion and TiO2. In
and enhanced charge separation from the substitution of Ti4+ with contrast, the samples prepared by the impregnation method
I5+ which led to the generation of titania surface states trapping experienced no shift, but absorption shoulders in the UV/Vis
electrons and suppressing recombination. The incorporation of spectra. This was caused by the creation of impurity energy levels
substitutional or interstitial metal dopants in the titania structure with the amount of metal ions dopant used determining their
generates trap levels in the band gap and thus modifying the band intensity.
gap after doping. As shown in Fig. 6, the trap levels usually in the The electronic structure of TiO2 doped with transition metals
form of narrow bands are located below the lower conduction band (Cr, Fe, Co, V, Ni and Mn) was also examined by Umebayashi et al.
edge. After modification, required energy level becomes hn (Eg [148] using ab initio band gap calculations based on density
 Et) where Et represents lower edge of the trap band level as functional theory. According to their work, a shift of the localized
opposed to hn Eg which is required for photon excitation before level to a lower energy was observed based on the increase of the
modification [134]. Consequently, electrons excited at these levels atomic number of the dopant. Incorporation of metal or its oxide
become trapped, with the holes having enough time for OH into TiO2 structure has been reported to cause an increase in the
generation such that electron/hole recombination is suppressed recombination rate between photogenerated electrons and holes
and overall process efficiency improved. The choice of metal via the impurity energy level. Therefore, doping can be effective if
dopant is determined by the ability of the metal to exhibit multiple the metal ions are placed near the photocatalyst surface where
oxidation states, possess ionic radii and Mn+/M(n+1) energy levels efficient charge transfer of the trapped electrons and holes can
closer to Ti4+ and the capacity to trap either electrons or holes. occur [41].
The type of metal dopant added will determine whether the In the field of CO2 photoreduction, Nie et al. [149] presented the
dominant charge carrier in the semiconductor will be either holes formation of smaller particles as a way by which doping can alter
in the valence band or electron in the conduction band [138]. Koci recombination rate. As smaller particles have large surface to
et al. [12] reported doping as a means of increasing the level of volume ratio, the migratory path is shorter such that the
holes in the band gap to permit the excitation of electrons where probability of the generated electrons and holes from the bulk
mobile holes are created in the valence band (p-type) or addition of undergoing recombination is reduced before reaching the surface.
an energy level fully occupied with electrons in the band gap which The dopant loading level plays a key role in CO2 photocatalytic
accelerates excitation into the conduction band (n-type). Carp et al. activity as increased product yield can be obtained due to red shift
[19] described n-type and p-type dopants where the former acts as towards visible light [144]. However, doping at high concen-
a donor centre of electrons and the latter conversely acts as trations results in the metal ions becoming recombination centres.
acceptor centers of holes. Recombination centers are formed in p- Gupta and Tripathi [41] further explained that increasing doping
type dopant as they have an affinity for hole formation once concentration results in a narrowed space charge layer where
negatively charged, while electron–hole recombination in n-type electron–hole pairs within this region can be efficiently separated
occurs due to increase in concentration of conduction electrons. by the electric field before recombination. However, exceeding the
optimum doping concentration results in an extremely narrow
space charge layer such that light penetration depth exceeds the
width of this space charge layer. Consequently, recombination rate
increases due to the lack of a driving force to separate them.
Koci et al. [150] used different loading ratios of Ag/TiO2 and
observed an increase in product yield of methane and methanol,
with 7% Ag/TiO2 showing the highest product yield compared to
lower loading ratios of 1%, 3% and 5%. Conversely, Sasirekha et al.
[82] obtained an optimal Ru loading value of 0.5 wt%, after which
photocatalytic activity decreased for 1.0 wt% due to increased
electron–hole recombination. Slamet et al. [20,86] also reported
that Cu dopant in excess of 3 wt% could reduce photocatalytic
activity by reducing the depth of light penetration, and thus
inhibiting interfacial charge transfer. When the doping content of
Fe3+ exceeded 0.03 wt%, Xin et al. [151] recorded a decrease in
photocatalytic activity due to electron–hole recombination, while
the opposite was observed for lower loading content (<0.03 wt%).
Regarding product formation, extensive studies into the use of
Fig. 6. Band structure of titania (a) before doping and (b) after doping where Eg,Ef,
doped TiO2 in CO2 photocatalysis have been conducted using
F, x, and s represent the band gap energy, Fermi level, work function, electron various metals such as chromium [152], copper [86,153], silver
affinity and semiconductor. Reprinted from Ref. [134] with permission. [154], platinum [91], palladium [155] and ruthenium [92]. Several
26 O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42

primary products with their yields includes methane


(4.8 mmol h1/gTi1 [91], methanol (2655 mmol/gcatal [86] and
4.12 mmol/gcatal [84] and formate (72.3 ppm [92]), respectively
have been obtained. However, thermal instability and increase in
recombination centers are drawbacks linked with this process
[24,156].

2.3.2. Metal semiconductor modification


The overall efficiency and surface properties of a semiconductor
can be altered by the addition of a metal which is not chemically
bonded to TiO2 [134]. These metals act as an electron scavenger and
thus facilitate the generation of holes. Ren and Valsaraj [157] and
Usubharatana et al. [13] described the addition of metals as a
source of charge-carrier traps which increased the life span of
separated electron–hole pairs and enhanced reaction rate. Fig. 7
illustrates the band structure of TiO2 before and after contact with
a metal. As shown in Fig. 7, the comparison of a semiconductor
Fig. 8. Metal modified semiconductor photocatalyst.
with work function Fs to a metal with work function Fm > Fs
results in the Fermi level of the semiconductor, EFs being higher
suppressed by the Schottky barrier of the metal in contact with
than the Fermi level of the metal EFm. When this metal is brought in
the surface of the semiconductor.
contact with TiO2 (Fig. 7b), electrons will flow from the
As a result of this trapping mechanism, the photogenerated
semiconductor to the metal until the two Fermi energy levels
electrons then diffuse to the surface of the adsorbed species where
reach an equilibrium. This results in an upward band bending
reduction takes place. Photoreactivity can be negatively influenced
formed due to an excess of positive charges in TiO2 generated from
by either a high concentration of metallic islands on the
the migrating electrons [158]. Consequently, this bending at the
semiconductor surface or an enhancement of their size [158].
metal–semiconductor interface creates a small barrier known as
When this occurs, reduced surface illumination of catalysts and
the Schottky barrier [32]. The Schottky barrier serves as an electron
increased recombination rate is observed. Krejcikova et al. [90]
trap which prevents migrating electrons from crossing back to the
conducted CO2 reduction studies using different loading ratios of
semiconductor and thus preventing recombination. The schematic
Ag/TiO2and observed increased product yield of methane in the gas
in Fig. 8 also illustrates the mechanism of a metal modified
phase and methanol in the liquid phase with increasing Ag
semiconductor for photocatalysis where recombination is
concentration under 254 nm UV irradiation over a 24 h period. The
increase in product yield compared to both commercial and
synthesized pure TiO2 was attributed to higher Fermi level of TiO2
and Schottky barrier formation which facilitated electron transfer
from the TiO2 conduction band to Ag particles and improved
charge separation, respectively. Tseng et al. [87] synthesized Cu
loaded TiO2 nanoparticles for the photocatalytic reduction of CO2
under UV irradiation using NaOH as a reductant. CH3OH
production was found to increase with increasing Cu concentra-
tion, after which a markedly decrease was observed when the
loading ratio exceeded 2 wt%. Optimum CH3OH production of
118 mmol/g was obtained using the 2 wt% Cu–TiO2 sample
following 6 h of UV illumination. The formation of the Schottky
barrier between Cu and TiO2 and the electric charge redistribution
via semiconductor–metal contact was reported to facilitate
electron trapping and thus promoting improved photo-efficiency.
Other than Schottky barrier, loading of noble metals such as Ag
[159], Cu and Au [160] on the TiO2 can enhance visible light
absorption via localized surface plasmonic resonance (SPR) effect.
This phenomenon can occur either by collective oscillation of
valence electrons in plasmonic nanostructures in resonance with
electric field part of inbound radiation or metallic elements
creating trap sites that propagates light within the semiconducting
material [161]. Morphology and size of the plasmonic nano-
structures influence the SPR frequency and intensity as well as the
resonant wavelength. Gas-phase photochemical reduction of CO2
using mesoporous TiO2 modified with bimetallic Au/Cu nano-
structures was studied for CH4 production under UV irradiation
[162]. The bimetallic nanocomposites were reported to exhibit
higher activity (CH4 yield of 11 mmol/gcatal) compared to Au/TiO2
and Cu/TiO2 (CH4 yield <4 mmol/gcatal).

Fig. 7. Band structure of titania (a) before contact and (b) after contact with a metal, 2.3.3. Non-metal modification
where the Schottky barrier is formed. (Fm and Efint represent the metal work Doping with non-metals creates heteroatomic surface struc-
function and Fermi level if titania is an intrinsic semiconductor respectively. tures and can modify the properties and activity of TiO2 toward
Reprinted from Ref. [134] with permission.
O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42 27

visible light [137]. Some of the non-metals that have been used 2.3.4. Co-doping
include nitrogen (N) [24], carbon (C) [118], sulphur (S) [163,164], The properties of TiO2 can also be modified via co-doping,
fluorine (F) [165]. Asahi et al. [24] and Asahi and Morikawa [156] which can be achieved via the combination of metal/metal, non-
described doping with anions as being more efficient for photo- metal/metal or non-metal/non-metal pairs. A synergistic effect can
catalytic activity compared to cations because they do not form be obtained with an appropriate combination of co-dopants
recombination centers caused by the presence of d states deep in compared to their single ion doped or undoped TiO2 [128]. During
the band gap of TiO2. Liu et al. [137] reported that effective band co-doping, the non-metal can cause a red shift in the visible light
gap narrowing can only occur by anion dopants if the non-metal region, while the metal can facilitate the transfer of photo-
has a comparable radius with O atoms and lower electro- generated charge carriers thus suppressing recombination. Apart
negativities than O, with the aim of facilitating uniform distribu- from co-dopants facilitating band gap narrowing, their combina-
tion and elevating the valence band. tion can result in the formation of different heterostructures (i.e.,
The band gap energy of TiO2 has been reported to be narrowed different electronic structures) with respect to TiO2 [44]. A
by a mixture of p states of the non-metal dopant with the O 2p heterostructure consisting of different combinations of non-metal
states of TiO2 via substitutional or interstitial doping [24,156]. and metal has the capacity for improved charge separation (metal)
Conversely, Valentin et al. [166,167] proposed that substitutional and visible light absorption (non-metal). Factors crucial for
doping with N results in the formation of localized levels within successful co-doping are the selection of the compatible co-
the band gap, with the catalyst synthesis conditions determining dopants and the method of introducing the dopants which affects
whether either interstitial or substitutional nitrogen exists in the the doping level [44]. Several metallic and non-metallic combi-
lattice of TiO2. On the other hand, Serpone et al. [168] attributed nations such as N–I [175], C–vanadium (V) [176] and Ag–V [177]
the origin of visible light absorption in their titania samples to the have been used in the photodegradation of pollutants and dyes.
existence of color centers instead of band gap narrowing via The results of these researchers showed the increased photo-
mixing of states, as proposed by Liu et al. [137]. catalytic activity and absorbance of visible light by the metal
First principle calculations by Asahi et al. [24] using anions (F, N, combinations compared to un-doped and single doped TiO2
P, S and C) indicated the superior activity of N owing to the p states systems.
influencing band gap narrowing through combination with O 2p For metallic combinations for CO2 reduction, the catalytic
states. Although S doping showed similar photoresponse as N, they activity of sol–gel derived Mn–Cu/TiO2 nanocomposites of varying
found the ionic radius of S too bulky to be integrated into the lattice metal concentrations was evaluated by Richardson et al. [178].
of TiO2. Zhang et al. [25] tested iodine doped TiO2 synthesized by After 24 h of UV irradiation, the photocatalytic activity of CO2 using
the hydrothermal method and found that the calcination 0.1 M NaOH and 0.25 M KHCO3 was found to be promoted based on
temperature influenced the rate of CO2 photoreduction under the coupling of Mn and Cu doped titania photocatalysts compared
visible light irradiation. They observed that increased calcination to either commercial TiO2–P25 or single metal loaded samples.
led to reduced surface area. An optimal yield of CO (2.4 mmol g1 Improved results were due to electron transport to the dopant
h1) was observed for the 10 wt% sample calcined at 375  C. Xue which suppressed electron/hole recombination. Maximum CH3OH
et al. [118] examined carbon doping for CO2 photoreduction using yield of 238.6 mmol/gcatal was achieved using the 0.22 wt%
citric acid as the carbon source and Na2SO3 as the reductant. After Mn/0.78 wt% Cu–TiO2 sample. The same trend was observed in a
6 h irradiation using high pressure 175W mercury lamp, further study conducted by Richardson et al. [179] using different
2610.98 mmol/gcatal of CH2O2 was produced. This was significantly sol–gel derived Cu–Ga/TiO2 nanocomposites of varying metal
higher than the undoped TiO2. concentrations. The photocatalytic activity was improved when Cu
Compared to other non-metals, N doped TiO2 (TiO2xNx) has and Ga doped photocatalysts were used such that the 0.78 wt%
been extensively studied because of its photoactivity toward Cu/0.22 wt% Ga–TiO2 sample gave the maximum HCHO yield of
visible light [137,169]. Zhao et al. [117] prepared N doped TiO2 394 mmol/gcatal when compared to single metal loaded samples or
nanotubes via hydrothermal method at different calcination TiO2–P25. They reported that their optimized results were due to
temperatures. N doping into TiO2 nanotube framework was found the rapid transfer of high energy electrons in their catalytic
to be effective for increasing the photoactivity of TiO2 in the visible structures.
light region compared to pure TiO2 and N doped TiO2. Optimum For metallic and non-metallic combinations in CO2 reduction,
total organic carbon content (sum of the product yields of co-doped N and Ni were introduced onto TiO2 framework for CO2
formaldehyde, methanol and formic acid) of 14,530 mmol/gcatal reduction using 0.2 mol/L of NaOH and Na2SO3 [115]. An increased
was observed using a N–TiO2 nanotube sample calcined at 500  C red shift toward the visible light was observed using the co-doped
for CO2 reduction with 0.1 N NaOH as reductant following 12 h of samples compared to pure titania and individually doped samples
light irradiation. of Ni–TiO2 and N–TiO2. An optimal methanol yield of
Since its quantum efficiency of anion doping is still low, 482 mmol/gcatal was observed after 8 h of UV light irradiation
investigations have been conducted by codoping with metals to using the 4 wt% N–6 wt% Ni/TiO2 sample compared to the
enhance the reaction rate [170,171]. Several transition metals such methanol yield of the individually doped samples of
as Pd, Fe and Pt have been used in the photodegradation of 245.4 mmol/gcatal of 4 wt% N–TiO2 and 214.4 mmol/gcatal of Ni–
pollutants and dyes [172–174]. The results of these researchers TiO2. They suggested that improved activity was due to the
showed the increased photocatalytic activity and absorbance of improved properties (surface area and crystallinity) of the co-
visible light by the metal ion modified TiO2xNx compared to bare doped samples and the synergy created by the metal (Ni) acting as
TiO2xNx. CO2 photoreduction studies have been carried out by an electron trap and the non-metal (N) facilitating increased
Varghese et al. [116] using N–TiO2 nanotube arrays with metals (Pt visible light absorption.
and Cu) under outdoor AM 1.5 sunlight. They found the optimal Li et al. [180] demonstrated that co-doped mesoporous Pt–N/
nitrogen concentration to be 0.75 atom% with Cu doping generat- TiO2 photocatalysts had some inherent advantages over undoped
ing greater hydrocarbon product yield of 104 ppm/ (cm2 h) TiO2. Using the optimum loading ratio of 0.2 wt% Pt for the
compared to Pt doping. As both Pt and Cu have varying effects synthesis of the co-doped samples under NH3 atmosphere, they
toward product selectivity; the combination of both metals observed increased CH4 evolution rate with increasing nitridation
resulted in an optimal yield of 111 ppm/cm2 h. temperature up to 525  C. After this temperature was exceeded; a
subsequent decrease in CH4 evolution was observed. Optimal CH4
28 O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42

production rate of ca. 2.6 mmol/gcatal1 was observed using the radicals being strongly absorbed on Ti sites due to the anions
0.2 wt% Pt–TiO2 sample when the amount of doped N was 0.84% on showing little solubility in these solvents of low polarity and
the basis of lattice oxygen atoms under visible light irradiation. therefore, CO was the major product observed during this reaction.
Improved activity under increasing nitridation temperature was When a high dielectric solvent such as H2O was used, CO2 – anion
due to the doping of more N atoms in the lattice position of oxygen radicals were greatly stabilized by the solvents which led to weak
in TiO2, which gave rise to improved visible light absorption. Above interactions with the surface of the photocatalyst, and thus
the optimum N doping concentration, decreased photocatalytic formate was observed, as the main product via the reaction of a
activity was observed due to increased defect sites and non- proton with the carbon atom of the CO2.– anion radical.
stoichiometry of the samples. Zhao et al. [187] employed titania supported cobalt phthalocy-
Co-doped samples were synthesized by Wang et al. [74] using anine (CoPc) nanoparticles for CO2 reduction in either NaOH or
commercial P25 TiO2 nanoparticles and CdSe quantum dots (QDs). Na2SO3 solutions. Maximum production of formic acid and
Pt was further incorporated by the wet impregnation methods formaldehyde was observed at concentration of 0.15 M due to
onto the CdSe–TiO2 samples for the experimental investigation for increased solubility of CO2 in NaOH and the OH ions produced
CO2 reduction using H2O. They found that the use of co-catalyst, Pt from NaOH acting as strong hole scavengers during the OH radical
with CdSe quantum dot (QD)-sensitized TiO2 heterostructures led formation. They further explained that electron/hole recombina-
to increased visible light absorption greater than their individual tion could be suppressed through the longer decay time of
photoresponse. No photocatalytic activity was also observed when electrons, since the holes are preoccupied in HCO3 formation in
either CdSe or Pt doped on TiO2 was employed for CO2 reduction. the CO2 saturated system. Further addition of an optimal
However, the synergy between CdSe–Pt/TiO2 heterostructures was concentration of Na2SO3 (0.1 M) led to an increase in formal acid
found to influence methane and methanol production under production through increased hole scavenging and proton
visible light irradiation with wavelength of 420 nm. In order to concentration within the semiconductor particle for CO2 reduc-
further demonstrate the need for heterostructure formation in co- tion. The same phenomenon was also observed by Tseng et al. [87]
doped samples, Wang and co-workers synthesized PbS–Cu/TiO2 in their CO2 reduction studies using NaOH solution.
samples with different sizes of quantum dots. Although they The study conducted by Koci et al. [188] demonstrated the
achieved optimal activity with the 4 nm co-doped heterostructure, influence of the volume of reductant on CO2 photocatalytic studies
the drawback of photocorrosion observed from oxidation in their using TiO2. The use of NaOH was reported to not only enhance CO2
previous and current studies could not be surmounted. Zhang et al. solubility, but also facilitate improved CO2 reduction via OH
[181] prepared co-doped Cu/I–TiO2 samples with different radical formation, which promoted the longer decay time of
concentrations using wet impregnation and hydrothermal meth- electrons. The production rate of CH4 and CH3OH was found to
ods. Under UV–vis and visible light irradiation, the photoactivity of increase when the volume of NaOH increased from 50 to 100 mL,
the co-modified sample was found to be higher than either of the and then markedly decreased above these values. For example, the
single ion modified catalysts (Cu–TiO2/I–TiO2). For CO production CH4 yield increased from 7.5 mmol/gcatal to >8 mmol/gcatal when the
under visible light, the optimum yield of 6.74 mmol/g1 was volume of NaOH increased from 50 to 100 mL, then decreased to
observed on the 1 wt% Cu–10 wt% I–TiO2 while the optimum yield <2 mmol/gcatal when 250 mL of NaOH was used. Ti-MCM-
of 12 mmol/g1 was observed on the 0.1 wt% Cu–10 wt% I–TiO2 41 mesoporous photocatalysts with Si/Ti molar ratios of 50,
sample under UV–vis light irradiation. The presence of the dopant 100 and 200 were tested for CO2 photoreduction using NaOH,
was reported to reduce the crystal size and influence visible light deionized H2O and monoethanolamine (MEA) as reducing agents.
absorption while Cu facilitated charge transfer and enhanced CO2 For the best photocatalyst within the series tested (Ti-MCM-
reduction. 41 with Si/Ti ratio of 50), maximum CH4 yield of 62.42 mmol/gcatal
was observed when MEA was used as a reductant compared to
3. Influence of operating parameters on CO2 reduction 5.62 mmol/gcata of CH4 over H2O after 8 h of UV illumination. The
lowest CH4 yield of 1.96 mmol/gcatal over NaOH was due to the
The following operating parameters listed below have been formation and precipitation of sodium bicarbonate (NaHCO3) in
shown to influence CO2 photoreduction: type of reductant, solution after contact with CO2 gas stream. Although the use of
temperature, pressure and particle size. These factors are discussed solvents other than water as hole scavengers can increase product
in the following sections. selectivity and yield, they still remain economically unsustainable
due to their potential to increase cost.
3.1. Effect of reductant
3.2. Effect of temperature
Several types of reducing agents such as H2O, NaOH, and
C3H7OH amongst others have been tested for CO2 photocatalytic CO2 photocatalysis is generally conducted at ambient con-
reduction [182–184]. For CO2 photoreduction using TiO2 to become ditions, i.e., room temperature because solubility decreases with
economically feasible, readily available sources of hydrogen are increasing temperature and the formation of electron/hole pairs
needed. H2O still remains the most naturally abundant source of occurs by photon (light energy) activation. An increase in reaction
hydrogen that is available and inexpensive [13]. Other reductants rate has been reported to occur at high temperatures due to
such as NH3, pure H2 gas etc., which serve as hydrogen sources are increased collision frequency and diffusion rate [188]. The
not readily available as primary feedstock and require prior optimum temperature required for photocatalysis is within
preparation [185]. However, the drawback of utilizing water is the 293–353 K, with decreasing activity occurring outside this range
low solubility of CO2 in H2O (2 g/L) and the competition of the CO2 [189]. This is due to exothermic adsorption of reactants being the
photoreduction process with hydrogen formation, as shown in Eqs. rate limiting step as the temperature approaches the boiling point
(4) and (6), which indicate that it is thermodynamically more of H2O. Yamashita et al. [190] demonstrated that photocatalytic
favorable to reduce H2O than CO2 [33,186]. reactions proceed more efficiently at temperatures higher than
Liu et al. [182] investigated the role of solvents on the product 275 K by using anchored titanium oxide catalysts for CO2
selectivity for CO2 reduction in an attempt to increase reaction reduction. They observed increased production rates of CH4, CO
yield. Their results indicated that the use of solvents with low and CH3OH under UV irradiation at 323 K compared to 275 K.
dielectric constants, such as CCCl4 and CH2Cl2, led to CO2– anion Saladin and Alxineit [191] studied the effect of temperature on CO2
O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42 29

reduction using titania samples irradiated under UV light for 4 h. Overall, the gross amount of hydrocarbons produced in the liquid
They found that the production rate of CH4 increased when the phase exceeded that of the gaseous phase. Similar trend of
temperature rises from 298 to 473 K. Based on model calculations, increased hydrocarbon production was observed when NaOH was
the reaction rate was not expected to be substantially improved used as the hole scavenger under CO2 pressurized conditions.
after the maximum temperature of 473 K due to hindered Additional reaction products such as C2H5OH and CH3CHO were
absorption of reactants. They also concluded that thermal also detected in the liquid phase. They suggested that improved
activation processes such as product desorption played a crucial reaction yield was due to the availability of CO2 on the surface of
role in improved reaction rate at 473 K. Product desorption readily TiO2, which accelerated CO2 reduction.
occurred at higher temperatures (473 K) compared to lower Koci et al. [188] demonstrated that the pressure influenced the
temperatures (298 K). rate of product yield obtained. They observed that the amount of
Guan et al. [49] investigated the use of a hybrid catalyst Pt CH3OH yield in the liquid phase increased when CO2 pressure was
loaded potassium hexatitanate (K2Ti6O13) combined with Fe based increased from 110 kPa (0.75 mmol/gcatal) to 130 kPa (1.5 mmol/
catalyst supported on Y zeolite (Fe–Cu–K/day) for CO2 reduction gcatal). Further pressure increase to 140 kPa led to reduced CH3OH
under concentrated sunlight. They found that the Pt/K2Ti6O13 yield. Conversely, CH4 yield in the gas phase increased with
catalyst produced H2 from water decomposition, while the increasing CO2 pressure from 120 kPa (2.75 mmol/gcatal) to 140 kPa
Fe–Cu–K/day catalyst reduced CO2 with the resulting H2 converted (4.5 mmol/gcatal). Kaneco et al. [186] studied the photocatalytic
into CH4, HCOOH and HCHO. The reaction temperature was found reduction of CO2 under various pressures using TiO2 suspended in
to promote the generation of the products listed above in addition iso-propyl alcohol medium. The results show that the pressure
to C2H5OH and CH3OH over the hybrid catalyst on temperature increase from 200 to 2800 kPa can increase CH4 production
increase from 534 to 590 K. They claimed that the simultaneous linearly. Conversely, the increased pressure conditions were found
supply of photons and thermal energy from the solar concentrator to inhibit CHOOH production, with reaction product observed at
was responsible for the optimal production of reaction products only 750 kPa. The lack of C2H4 formation was attributed to the
observed at 590 K. The same phenomenon with higher production accelerated formation of CH4. An increase in CH3OH formation
rate was also observed by Guan et al. [192] when they used a hybrid from 175 to 230 mmol/g-cat was observed by Tseng et al. [87] when
catalyst Pt loaded potassium hexatitanate (K2Ti6O13) combined the CO2 pressure increased from 110 kPa to 125 kPa. Further
with Cu/ZnO catalyst under concentrated sunlight at 583 K. pressure increase above 125 kPa led to decreased CH3OH produc-
Increased methanol yield on temperature increase within the tion rate of 85 mmol/g-cat. Cost of fabricating sophisticated high
range of 333–373 K was also observed by other researchers during pressure systems must be considered when selecting parameters
CO2 reduction studies [20,193]. for reactor designs in CO2 reduction.
Photoreduction studies conducted by Kaneco et al. [194]
demonstrated that temperature had no effect on the catalytic 3.4. Effect of particle size
activity of their samples. The photocatalytic activity of TiO2
suspended in supercritical CO2 was investigated. Formic acid Particle size is a key parameter in photocatalytic processes
production observed in the liquid phase was attributed to the since the interaction between the amount of absorbed and
reaction of water with reaction intermediates on the surface of TiO2. reflected photons and the reactants depends on it. Apart from
An increase in temperature at the rate of 278 K from 308 to 323 K led nanostructured photocatalyst possessing high surface area, they
to the steady state formation of formic acid. Koci et al. [188] reported also benefit from low refractive index which minimizes light
that a temperature increase of 10 K from 299 K to 309 K did not reflection, high surface to volume ratio and rapid charge transfer
influence the hydrocarbon production rate for the photocatalytic [196,197]. Several researchers have established that photo-
reduction of CO2 using TiO2 following 4 h of UV irradiation. Although catalysts in the form of nanoparticles are more effective than
increasing reaction temperatures have been reported to facilitate bulk powders [198–200]. The rate of electron–hole recombina-
increased production rates, the cost of fabricating sophisticated high tion has been reported to be controlled by particle size since
temperature photoreactor systems capable of maintaining the extremely small ultrafine particle (within the diameter range of
selected optimum temperatures and the source of thermal energy few nanometres) experience surface recombination as opposed
required to heat up solvents i.e., water has a high specific heat to large particles where volume recombination predominates
capacity, still poses a problem. [200]. The problem of volume recombination can be overcome
by reducing the particle size. During surface recombination,
3.3. Effect of pressure most of electron–hole pairs photogenerated close to the surface
undergo rapid recombination due to their shorter migratory
Improved product selectivity has been reported to occur due to paths, abundant surface trapping sites and limited driving force
increased CO2 concentration resulting from an increase in CO2 for charge separation [138]. This phenomenon has been reported
pressure in aqueous media [188]. Experimental studies conducted to occur within certain size reduction. Zhang et al. [200] further
by Mizuno et al. [195] demonstrated the effect of CO2 pressure on demonstrated that the number of available active surface sites
CO2 reduction using TiO2 suspensions in H2O and NaOH. CH4, C2H4 and transfer rate of surface charge carrier increased with smaller
and C2H6 were observed in the gas phase under pressurized particle sizes due to their larger surface area. They observed
conditions (2500 kPa), with no hydrocarbon production detected increased photoactivity in the decomposition of CHCl3 when
under ambient pressure. CH4 production was found to increase TiO2 particle size decreased from 21 to 11 nm. Decreased
sharply with increased CO2 pressure from 500 to 2500 kPa. Slight photoactivity was also observed when the particle was further
increase in formic acid production was observed in the liquid phase reduced to 6 nm. Optimal photoactivity was demonstrated with
under similar pressurized conditions. Overall, increased CO2 the 10 nm particle. On the other hand, Koci et al. [198] proposed
pressure accelerated CO2 reduction when both H2O and NaOH that the 14 nm TiO2 particle was the optimal value for CO2
were used. reduction, since they obtained maximum CH4 and CH3OH
On the other hand, CHOOH and CH3OH production was production using this particle size. A decrease was observed
observed in the liquid phase at ambient pressure. While a linear on further increase to 29 nm. They attributed the decreased
increase of formic acid was observed with slight pressure increase, photoactivity observed in samples with particle size <14 nm to
CH3OH yield was found to reach an optimal rate at 1000 kPa. rapid flocculation which decreased availability of active sites.
30 O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42

The model developed by Almquist and Biswas [199] was used the catalyst film which is connected to an external potential that
to elucidate the effect of particle size on TiO2 activity for the can move excited electrons, and thereby, reducing electron–hole
photo-oxidation of phenol. Particle sizes ranging from 5 to recombination to improve efficiency.
165 nm were prepared from flame synthesized TiO2 and Highly dispersed titanium oxide anchored onto Vycor glass
commercially available P25 and anatase TiO2. The results also was tested for the photocatalytic reduction of CO2 with H2O [190].
highlighted the strong dependence of particle size on photo- The supports were prepared through facile reaction between
activity. Increased photo-oxidation occurred when the particle surface OH groups of a transparent porous Vycor glass and TiCl4.
size increased from 5 to 30 nm and decreased when the particle UV irradiation of the support led to the formation of C1
size increased beyond this value. The optimal particle size range compounds such as CH4, CH3OH and CO as major products and
reported was within the range of 25–40 nm. Band gap of trace amounts of C2 compounds (C2H4 and C2H6) at 323 K. Cu
photocatalysts has been reported to be influenced by particle nanoparticles were deposited on transparent conductive fluori-
size [41,138,197]. They proposed that semiconductor particles nated tin oxide (FTO) glass substrates for CO2 reduction to CH4
within the nanometre range experienced energy shift according under UV irradiation [212]. Cu–TiO2 films were reported to
to the size quantum effect which could accelerate reduction and exhibit higher yields compared to pure TiO2 and TiO2 P25.
oxidation reactions via the conduction and valence band, Enhanced light absorption and increased diffusion length of
respectively. This size quantization effect expected to cause an photoinduced electrons were amongst some of the reasons for
increase in the band gap energy results in a shift to larger redox enhanced CO2 photoconversion rates. TiO2 pellets (Aerolyst 7708)
potentials, which increases rate constants for surface charge were affixed to a flat glass tray by Tan et al. [83,89,213] to increase
transfer. Banerjee [201] also reported that the large fraction of absorption capacity and contact area for CO2 photoreduction. The
surface atoms and high surface to volume ratio found in product yield of CH4 (200 ppm) using ultraviolet light C (UVC)
nanoparticles are responsible for enhanced light absorption wavelength of 253.7 nm was reduced to values lower than
through indirect electron transition at the boundary of the crystal 100 ppm on switching to ultraviolet light A (UVA) wavelength of
i.e., surface or interface between two crystals. Particle size of 365 nm after 48 h of irradiation.
photocatalysts must be carefully considered during catalyst Platinum (Pt)–TiO2 nanostructured thin films with different
synthesis since it is invariably linked to surface area and deposition times were prepared by Wang et al. [213] for
photocatalytic efficiency. immobilized onto indium tin oxide (ITO)-coated aluminumosi-
licate glass using RF magnetron sputtering and gas-phase
4. Catalyst configuration: supports deposition method. The films which had a one-dimensional
structure of TiO2 single crystals with ultrafine Pt nanoparticles
TiO2 can be synthesized as powders (nanospheres, micro- (NPs, 0.5–2 nm) were found to exhibit enhanced CO2 photore-
spheres), crystals, films or immobilized by dip or spin coating onto duction efficiency with selective CH4 yield of 1361 mmol/gcatal h.
substrates such as fibers, membranes, glass [202], monolithic The fast electron-transfer rate in TiO2 single crystals and the
ceramics [155], silica [203] and clays such as zeolite [204], efficient electron–hole separation by the Pt NPs were the main
kaolinite [95], montmorillonite [205] etc. Several materials have reasons reported to be attributable for this enhancement.
been used as TiO2 support for CO2 reduction. The use of supports Mesoporous Cu–TiO2 nanocomposites synthesized by a one-pot
eliminates the need for post treatment separation, provides high sol–gel method were loaded onto glass fiber filters as thick films
surface area and mass transfer rate [202]. The product selectivity, for CO2 photoreduction to CO and CH4 [94]. CH4 and CO peak
structure and electronic properties of TiO2 can be modified by the production rates of 10 and 60 mmol/gcatal h were achieved over
use of supports. However, the photocatalyst must be strongly the 0.5% Cu/TiO2–SiO2 composite. Improved results were reported
adhered to the support and have light absorption properties to be to be influenced by the synergistic effect resulting from the
effective. An ideal support must be resistant to degradation combination of the SiO2 substrate and Cu deposition loaded onto
induced by the immobilization technique and should provide firm the glass fiber filter.
adhesion between the support and the catalyst [206,207]. Mass The effect of Ag/TiO2 nanoparticles deposited on glass
transfer limitations and low light utilization efficiency due to little microfiber filter for UV light induced CO2 photoreduction using
or no light absorption in the pores or channels of the catalyst water vapor as electron donor was performed by Collado et al.
coated supports are key limitations that have been identified with [98]. Deposition of Ag on TiO2 surface led to an enhancement in
the use of supports [208]. Many researchers have focused on ways the production of C1–C3 compounds which increased as Ag
of anchoring photocatalysts onto supports since high photo- loading increased from 1.5–3.0 wt%. Better catalytic performance
conversion efficiencies and improved light harvesting can only be and selectivity were observed over glass filters containing Ag
achieved through the combined use of optimized photoreactor and samples prepared by wet impregnation than incipient wetness
photocatalyst configurations. An overview of some commonly impregnation procedure. Enhanced hydrocarbon production was
used supports is presented below, including glass, optical fibers reported to be due to lower recombination rates and synergistic
and monoliths. effect between TiO2 and Ag nanoparticles. The transparency of the
glass material used can also limit the overall efficiency due to the
4.1. Glass catalyst receiving insufficient light e.g., Pyrex glass can cut off UV
light below 300 nm [214]. On the other hand, quartz glass is a
Several types of glass substrates such as beads [209], plates better alternative as a light conducting material because of its
[210], microfiber filter [94,98,211] and plates [102] have been used excellent light transmission properties and its ability for
for CO2 reduction due to the transparency of the substrates to light increased contact efficiency, thus creating more active sites.
irradiation. The use of conductive materials like glass as supports
have been extensively studied due to their ability to prevent total 4.2. Optical fibers
internal reflection through surface roughening which also provides
better catalyst adhesion to the glass substrate and increases the The use of a single or bundle of optical fibers for the remote
amount of immobilized catalyst per unit area [28]. Furthermore, delivery of light to reactive sites of coated photocatalysts has been
Ray and Beenackers [28] reported that utilizing conductive studied by several researchers for wastewater treatment and CO2
materials serve as a means by which light can be transmitted to photocatalysis [48,154,215–217]. All researchers observed
O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42 31

increased degradation and conversion rates when the optical fibers 4.3. Monoliths
were simultaneously used as a support and light distributing
guide. In optical fiber, light is transmitted along the fiber core by The use of interconnected three-dimensional structures like
the cladding with lower refractive index that traps light in the core honeycomb monoliths containing parallel straight channels has
through total internal reflection. Light can be primarily emitted at been exploited for industrial processes due to its potentially high
the end of the fiber (end emitting) or through the leakage of light as surface to volume ratio, easy of scale-up through an increase of its
it travels from the fiber core to the cladding via the side surfaces dimensions and channels, control of structural parameters (i.e.,
(side emitting) [218]. pore volume, pore size and surface area) etc. [223,224]. Different
Catalytic performance of 120 catalyst coated optical fibers types of metal oxides and mesoporous materials have been
was evaluated for CO2 reduction under 365 nm UV irradiation. immobilized on TiO2 coated monolithic materials to improve
Maximum CH3OH yield of 0.45 mmol/gcatal h was achieved using catalytic performance for CO2 photoreduction. Tahir and Amin
1.2 wt% Cu–TiO2 catalyst coated fibers. Properties of the optical [225] deposited montmorillonite (MMT) based TiO2 onto
fibers such as external surface area and light transmittance were monolithic structure to improve surface area and adsorption of
reported to influence the processing capacity of the catalysts. gaseous species for CO2 photoreduction with H2O. The addition of
The influence of the optical fibers in delivering photons required MMT into TiO2 matrix was reported to increase surface area from
to activate different catalyst combinations such as Cu–Fe/TiO2, 42.98 m2/g for pure TiO2 to 51.79 m2/g for MMT/TiO2. Higher CO
Cu–Fe/TiO2–SiO2 and N3 dye–Cu–Fe/TiO2 was further demon- (52 mmol/gcatal1 h1) and CH4 (139 mmol/gcatal1 h1) production
strated [48,71,219]. Overall, maximum CH4 production rate of was achieved over the MMT/TiO2 monolith compared to pure TiO2
1.86 mmol/gcatal h was obtained using Cu (0.5 wt%)–Fe (0.5 wt (CO, 47 mmol/gcatal1 h1 and CH4, 82 mmol/gcatal1 h1) after 10 h
%)/TiO2–SiO2 catalyst coated fibers while ethylene production of UV light irradiation. Photocatalytic activity was reported to be
rate of 0.575 mmol/gcatal h was achieved with Cu (0.5 wt%)–Fe influenced by increased CO2 adsorption originating from surface
(0.5 wt%)/TiO2 coated fibers under UVA irradiation. During gas hydroxyl (OH) groups in MMT/TiO2 framework. The effect of
phase CO2 reduction studies, Wang et al. [193] obtained yields of monolithic geometry such as cell density and channel length on
11.3 mmol/gcatal h and 11.3 mmol/gcatal h for methanol production UV induced CO2 photocatalysis was further evaluated. The
under visible light irradiation and sunlight, respectively, when product rates were reported to be influenced by the geometry
the optical fibers were coated with NiO/InTaO4. of the monolith since maximum product rates of CO and CH4 were
32 optical fibers coated with inverse opal Cu–TiO2 inserted in a observed over the monoliths with lower cell density of 100 cells
stainless tube were tested for the photoreduction of CO2 to CH3OH per square inch (cpsi) and channel length of 2.5 cm compared to
[157]. CH3OH product rate of 0.036 mmol/gcatal h was achieved the monolith with higher cell density (400 cpsi) and lengths of
after UV irradiation with light intensity of 113.65 mW/cm2. The 1.2, 1.7 and 5 cm. The monolith channel length was reported to be
inverse opal configuration was reported to enhance catalyst linked to light distribution. The effect of In loading on TiO2 coated
activation via increased contact time of light within the photo- monoliths was further tested for photocatalytic reduction of CO2
catalyst. Although high catalyst loading and direct light excitation under UV irradiation by the same research group [226]. The
of coated catalyst films can be achieved when a bundle of optical introduction of indium to TiO2 framework not only increased
fibers are coated, fragility of the optical fibers and the durability of surface area and reduced particle size, but also facilitated charge
their coatings has been described as drawbacks associated with transfer. Maximum CH4 production rate of 55.4 mmol g1 h1 was
their use [197]. The durability and performance of these fibers are observed over 10 wt% In/TiO2 monolith with 100 cpsi after 10 h of
directly related to the adhesion of the catalyst coatings on the UV irradiation.
fibers and thickness of the coated layer which may not withstand Photocatalytic studies conducted using monoliths as catalytic
severe gas/liquid continuous flow conditions [220,221]. Even support for wastewater treatment, NO and CO2 reduction have
though roughening of the fiber surface has been reported to identified low light utilization efficiency due to little or no light
increase durability of these coatings, the problem of uneven absorption in the pores or channels of the honeycomb monolith
catalyst and light distribution has also emerged [220,221]. Heat as a major drawback associated with its use [155,208,223]. Not all
build-up from the bundled array of fibers can result in catalyst immobilized photocatalyst may be activated due to limited light
deactivation [222]. distribution arising from the catalyst coated on the outer surface
absorbing most of the light and its intensity decaying rapidly
along the opaque channels of the monolith [207,227]. In a

Fig. 9. Schematic of light propagation in a single channel of a coated honeycomb monolith threaded with a non-coated side light emitting optical fiber.
32 O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42

mathematical model developed by Hossain et al. [228] for influx observed for the stainless steel support coated with 45% CdS–TiO2
of UV light within a square channel monolith, half of the incoming nanocomposite.
light flux was reported to be lost due to light shadowing effect at The direct conversion of CO2 over TiO2 coated stainless steel
the entrance of the channel of the monolith wall. The UV light flux webnets of varying sizes was investigated under UV irradiation
was also reported to decrease sharply with increasing distance in [231]. High surface area and good utilization of UV light were
the monolith channel. A strategy for improving light distribution observed on TiO2 films deposited on the webnets. An increase in
in monolithic structures was originally proposed by Du et al. mesh size resulted in increased TiO2 surface area and reduced
[220], where non coated side-light emitting fibers were evenly penetration of UV light. Evaluation of the photocatalytic activities
distributed in each TiO2 coated channel to ensure light refracted of TiO2 coated on three different mesh sizes of stainless steel
out of the surface of the fiber could reach the catalyst–reactant webnets for CO2 photoreduction resulted in higher product rate
interface without attenuation. Fig. 9 shows the schematic where for TiO2 coated on 120 mesh size than TiO2 coatings on 60 and 200
the gaseous reactants diffuse into each coated interconnected mesh sizes. Nishimura et al. [232] dip coated TiO2 on a silica–
monolithic channel, adsorb and react with catalyst activated by a alumina gas separation membrane to obtain 3.5 ppmV/h of CO
non-coated side-light fiber to form desorbed products and after 336 h, while Pathak et al. [233] used the hydrophilic
intermediates which diffuse back into the bulk gas stream (CO2 structural cavities in Nafion-117 membrane films to host TiO2
saturated with H2O). coated with nanoscale silver and obtained 0.071 mg1 and
In recent years, studies on CO2 reduction using non coated 0.031 mg1 of methanol and formic acid after 5 h. Reproducible
side-light emitting fibers with geometric notches in the core– results were obtained when these films were reused. Cybula et al.
cladding system were reported to improve photocatalytic activity [234] employed a flat perforated steel or plastic tray as a support
[229,230]. Vapor phase CO2 with H2O was reduced to CH3OH by for the dispersion of TiO2 in a tubular reactor designed for CO2
NiO/InTaO4 coated monoliths containing no fibers, bare fibers and photoreduction studies. They observed that the type of support
fibers with tip-reflection and mid-carves under visible light used not only played a critical role in determining the amount of
irradiation. Highest CH3OH rate of 0.16 mmol g1 h1 was immobilized catalyst but also influenced the photoconversion
observed over 1 wt% NiO/InTaO4 monolith containing fibers with rate when the same coating procedure was used. A decrease in
tip-reflection and mid-carves compared to the monolith con- catalyst loading and methane production (from 90 ppm to
taining no fibers and bare fibers. The results are linked with fibers 34 ppm) was observed when the support was switched from
with tip-reflection and mid-carves having the highest side light steel to plastic due to weaker adhesion of TiO2 on plastic
emission percentage of 98% compared to no fiber (84%) and bare compared to steel. CO2 photoreduction studies by Shioya et al.
fiber (93.8%) amongst configurations. Overall, maximum acetal- [203] and Li et al. [94] employed silica as supports due to its even
dehyde conversion rate of 0.3 mmol g1 h1 was achieved with the composition and orderly mesoporous structure with small
2.6 wt% NiO/InTaO4 monolith containing fibers with tip-reflection channels. Sasirekha et al. [82], Yang et al. [93] and Li et al. [94]
and mid-carves by simulated sunlight AM1.5G at 70  C. The attempted to improve this arrangement by doping with metals
loading of different metal-based TiO2 nanomaterials onto such as Ru and Cu. They found that the combination of metals
monolithic structures threaded with optical fibers were tested with mesoporous silica enhanced the reaction rate due to
for UV and visible light induced CO2 photocatalysis [102,230]. The effective TiO2 dispersion and improved absorption of CO2 and
loading of metal or metal oxide on the surface of TiO2 via the H2O on the surface of SiO2. Product yields of 60 mmol g–
introduction of defects into the lattice was reported to tailor its TiO21 h1 and 10 mmol g–TiO21 h1 were obtained for CH4
band width towards the visible light and alter its particle and CO, respectively, in a continuous flow photoreactor at an
properties. For example, sol–gel derived 1 wt% Ni2+-based TiO2 optimal doping ratio of 0.5 wt% Cu using Xe arc light source [94].
monolith containing optical fibers showed improved CH3OH Maximum CH4 production of 627 mmol g1 h1 was observed by
production rate of 13 mmol/gcatal h compared to pure TiO2 coated Yang et al. [93] with 2 wt% Cu after 8 h reaction time. Improved
monolith. Addition of Ni2+ influenced activity and selectivity of surface area and better dispersion of cerium (Ce)–TiO2 on
TiO2 toward UV and visible light region due to the substitutional mesoporous silica (SBA-15) was also demonstrated by Zhao
metal ions not only causing changes in the electronic structure et al. [97] in their CO2 photoreduction studies following 4 h of
and light absorption properties of TiO2, but also altering the UV–vis irradiation. They found that an optimal amount of 3%
surface area, grain size and degree of phase transformation. Ce–TiO2 dispersed on the silica matrix (Ti:Si—1:4) not only
facilitated improved textural properties compared to pure TiO2,
4.4. Other supports but also resulted in an order of magnitude increase in CO
(7.5 mmol g1) and CH4 (7.9 mmol g1) production. They reported
The effect of co-catalyst (Cu–Pt)-sensitized TiO2 nanoparticle that the adsorption properties of silica resulting from its unique
wafer on CO2 photocatalytic conversion was studied under full mesoporous structure was one of the contributing factors due to
sun illumination (AM 1.5G) [103]. Coated wafers which had the increased localized CO2 concentration near TiO2 surface
randomly connected pores were used as flow through membrane where photocatalysis could occur. Clays have been extensively
such that reactants (CO2 and H2O) pass through one end of the used as supports in photocatalytic studies because of their low-
membrane and products collected at the other end. Improved cost and strong absorption capacity [235]. Koci et al. [95] used
product rates were achieved due to back reactions limited by the kaolinite/TiO2 in CO2 photoreduction and obtained CH4 and
diffusion of reaction products to the outlet. Optimum amount of CH3OH yields of 4.5 mmol/gcatal and 2.5 mmol/gcatal after 24 h of
catalyst loading on the TiO2 wafer were 9.6 at% Cu and 8.7 at% Pt. irradiation. Kaolinite prevented particle aggregation and modi-
Hence, maximum conversion of CO2 to CH4 (49 ppm/cm2 h) was fied the acid–basic properties of the surface of TiO2. The use of
achieved over TiO2 wafer sputtered with both Cu and Pt layers montmorillonite as support in CO2 photoreduction has also been
than Pt (28 ppm/cm2 h) and Cu (38 ppm/cm2 h). Thin layer of examined by Kozak et al. [205]. CH4, CH3OH and CO production
CdS–TiO2 nanocomposite was coated on a stainless steel support were observed over ZnS after 24 h of irradiation. Carbon based
to improve CO2 reduction performance. Performance of catalyst materials such as graphene/graphene oxide [236–239], carbon
coated supports was dependent on metal concentration and size nanotubes (CNT) [240] and fullerenes amongst others have
of the nanoparticles. Maximum CO and CH4 production rates of attracted wide attention as support materials for TiO2 induced
10.5 and 1.5 mmol/gcatal under visible light irradiation were CO2 photocatalysis, due to their high specific surface area,
O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42 33

electronic properties and enhanced transport of photogenerated procedure [251]. Watson et al. [64] demonstrated the preparation
electrons and visible light absorption [241–244]. Recently, hollow of more uniform and pure photocatalysts via the alkoxide route,
spheres consisting of alternating titania (Ti0.91O2) and graphene while Sivakumar et al. [252] used ammonium nitrate and titanyl
nanosheets were tested for CO2 reduction to CO and CH4 in the sulphate as precursors. The rapid hydrolysis rate of titanium alkoxide
presence of water vapor [243]. CH4 and CO production rates over has been reported as a major drawback that makes this process
Ti0.91O2–graphene were 1.14 and 8.91 mmol/g h, respectively, difficult to control [253]. The sol–gel process is initiated via
which was reported to be five times higher than Ti0.91O2. Multi hydrolysis and polycondensation of metal precursors
walled CNT (MWCNT)/TiO2 nanocomposites were reported to (Eqs. (21)–(25)) where R stands for C4H9 [254,255].
exhibit superior photocatalytic activity for CO2 photoreduction Apart from esterification, the hydrolysing water can also be
compared to anatase TiO2 and pure MWCNT [244]. Maximum CH4 introduced and controlled through oxolation, as shown in Eq. (22).
yield of 0.178.91 mmol/gcatal h was achieved after 6 h of visible During condensation, the crystal of the metal oxide can be formed
light irradiation. Thus, the choice of an adequate support is of when the constituent particles of the gel are pulled into a compact
utmost importance since the overall process efficiency of the mass. Additionally, acetic acid modifies and stabilizes the hydrolysis
photoreactor is predominantly determined by the amount of process by altering the alkoxide precursor at molecular level and
activated photocatalysts. It is therefore imperative to utilize acting as a chelating ligand, such that TiOH bonds were formed
versatile materials with excellent light transmission properties when the bidental ligand was broken off (Eq. (25)). The decrease in
that can simultaneously serve as catalyst carrier and provide high the hydrolysis rate results in the formation of fine particles of titania
light transfer area via light distribution from the source to the which are uniformly dispersed in solution. Appropriate amount of
photocatalyst present within the photoreactor. metal precursor(s) can also be introduced within the hydrolysis and
polycondensation phase depending on the weight percent loading
5. Support immobilization techniques calculated from the amount of precursors used in the procedure. Sols
are usually deposited on the substrates via dip-coating, spin coating,
TiO2 based catalysts can be deposited on structured substrates doctor blade techniques amongst others. The withdrawal speed of
through aqueous or gaseous routes. Some examples of aqueous the substrate, number of coating cycles and the sol viscosity
methods include sol–gel and electrophoretic deposition, while gas determines the catalyst film thickness. TiO2 was immobilized onto
phase methods include spray pyrolysis deposition, chemical vapor gas separation membrane by sol–gel dip coating method to study the
deposition and physical vapor deposition. Table 3 highlights the CO2 reduction performance [256]. The gas separation membranes
advantages and disadvantages of different methods used for were dipped into TiO2 sol with different withdrawal speeds.
immobilising TiO2 catalysts [245–247]. Maximum CO production yield of >250 ppm V was observed over
the TiO2 film coated with the withdrawal speed of 0.66 mm/s after
5.1. Sol–gel method 72 h of UV irradiation. The improved activity using this optimum
coating condition was attributed to TiO2 films being thin and even.
Sol–gel technique is amongst the most widely used procedure for The synthesized nano-sized TiO2 films have been to have special
preparing TiO2 photocatalysts. This technique is not only noteworthy catalytic properties due to the integration of the active metal during
for achieving excellent chemical homogeneity but, also deriving the gelation stage.
unique stable structures at low temperatures as well [30,248–250].
The compositional and microstructural properties of the nano-sized & Esterification:
samples can be tailored through the control of the precursor
ROH + CH3COOH H2O + RCOOC2H5 (21)
chemistry and processing conditions. Inorganic metal salts like
titanyl sulphate, titanium tetrachloride etc., (non-alkoxide) and
metal alkoxides e.g., titanium(IV) butoxide are usually employed as
chemical precursors. Conversion from the liquid sol phase into the & Oxolation:
solid gel phase occurs due to solvent loss and complete polymeriza-
Ti(OR)3 (OH) + Ti(OR)3 (OH) ! (RO)3 TiOTi(OR)3 + H2O (22)
tion. The pH of the reaction medium, water:alkoxide ratio and
reaction temperatures are factors that influence the sol–gel

Table 3
Advantages and disadvantages of different methods used for immobilizing TiO2 photocatalysts.

Catalyst preparation Advantages Disadvantages References


method
Sol–gel High purity of materials Hydrolysis rate is difficult to control [248–251]
Homogeneity Longer processing time compared to CVD and PVD
Versatile means of processing and control of parameters Calcination at high temperatures results in the decomposition of
substrates
Large surface area Multiple coating cycles is required depending on the sol viscosity
Chemical bonding results in strong adherence of coating to
the substrate

Physical vapor Does not require complex chemical reactions Expensive as vacuum systems are used [19,245]
deposition Low to medium deposition temperature Difficulty in deposition of multiple source precursors due to various
evaporation times

Chemical vapor Suitable for uniform and pure films with high deposition Could be expensive if vacuum systems are used [245–247]
deposition rate
Short processing time Difficulty in deposition of multiple source precursors due to various
evaporation times
High deposition temperature (>600  C) is required
34 O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42

& Hydrolysis: particle aggregation, which affects the microstructure and textural
properties (crystallinity, surface area, morphology) of TiO2
Ti(OR)4 + H2O ! Ti(OR)3 (OH) + ROH (23)
[41,255]. The removal of organics has been reported to occur at
temperatures 673 K [254]. Conversely, well crystallized anatase
TiO2 films can be synthesized by the peptization of tetraisopropyl
& Condensation dehydration: orthotitanate in acidic conditions at low temperatures [257].
Since photoconversion can only occur when the photogen-
Ti(OR)3 (OH) + Ti(OR)4 ! (RO)3 TiOTi(OR)3 + ROH (24)
erated holes and electrons are present on TiO2 surface, the surface
phase of TiO2 exposed to reactant and light has been found to play a
critical role in determining the rate of photoconversion by several
& Chelation: researchers [62,75,258]. Consequently, a reasonable calcination
temperature must be selected such that increased crystallinity is
H R achieved with the surface area remaining intact and unchanged. Su
et al. [255], Vijayan et al. [259] and Schulte et al. [260]
O
demonstrated that their optimum activities were strongly depen-
RO OR dent on the crystallinity of their nanostructures. Su et al. [255]
showed that optimal photocatalytic activity for decomposition of
Ti salicylic acid can be achieved with the sample calcined at 773 K.
RO The photoactivity decrease was observed beyond this temperature
OR
due to reduced surface area and decreasing anatase fraction. The
O anatase to rutile fraction was found to decrease with increasing
temperatures.
H R Conversely, Vijayan et al. [259] observed increased methane
production using titania nanotubes calcined between 473–673 K
-4 ROH + 2CH3COOH for CO2 reduction owing to the combined effects of crystallinity
and surface area. Declined activity was observed on further
R temperature increase. The anatase content was also found to play a
critical role in the UV activation of nanotubes prepared by Schulte
O et al. [260] for CO2 reduction. They also observed declined
O O reactivity with increasing calcination temperature. Increased rutile
content at near 953 K was found to tune the photoresponse toward
H3C C Ti C CH3 the visible light region which further optimized reactivity of the
O O samples. Amongst the crystalline phases of titania that can be
O formed during calcination, anatase has been reported to possess
better photocatalytic activity for CO2 reduction compared to rutile
due to higher surface areas and improved hole trapping arising
R
from steeper band bending [251]. Phase transformation from
(25) anatase to stable rutile can occur upon thermal treatment between
The series of Cr doped TiO2 samples synthesized by the sol–gel temperatures of 623–1373 K owing to different processing
method was found to promote the CO2 reforming performance of methods, precursors and techniques of determining this transition
TiO2. Under their experimental conditions, the Cr-doped samples temperature [143,261]. The presence of mixed crystalline phases of
showed improved photoresponse in the visible light in their study titania (i.e., anatase and rutile) has also been reported to show
compared to the pure TiO2 film. Optimum product yields of improved photocatalytic activity due to synergistic effect derived
92.5 mmol/gcatal (CO), 15.1 mmol/gcatal (CH4) and 19.1 mmol/gcatal from better charge separation and high surface area [262].
(C2H6) were obtained using the 7 wt% sample. Copper supported Improved charge separation and high reactivity at the anatase
on unstructured and inverse opal titania (templates of polystyrene to rutile interface occurs during electron transfer from rutile to
spheres) films coated onto optical fibers by sol–gel dip-coating anatase at this interface where defect sites with unique charge
technique were employed for the photoreduction of gaseous CO2 to trapping and adsorption properties can be created [62,262,263].
CH3OH in the presence of water vapor and UV light [157]. Although Zhang et al. [258] further investigated the relationship between
the methanol production rates of the Cu films supported on inverse the effect of calcination temperature and time on the surface
opal titania (0.0364 mmol/gcatal h) were comparable to the films phases and photocatalytic activities of TiO2. The transformation
supported on unstructured titania, much higher quantum effi- from anatase bulk phase to rutile occurred at 823 K, with the
ciency was achieved using the inverse opal film due to improved anatase phase still being maintained on the surface till 680  C.
photon utilization rate observed at a lower light intensity. Further temperature increase to 973 K led to the complete
conversion of the bulk anatase phase into rutile, with only 44% of
5.1.1. Thermal treatment the anatase phase being present on the TiO2 surface. Maximum
Calcination is one of the means by which crystal growth can be hydrogen production was observed for samples calcined between
controlled. The crystal growth influences the phase, shape, size and 973–1023 K due to catalyst's surface consisting of a mixed phase of
surface area of photocatalysts. Sol–gel derived precipitates tend to anatase and rutile particle, with the bulk consisting of almost pure
be amorphous in nature and require heat treatment to remove rutile. Similar results were also observed for the samples calcined
organic molecules from the final products and induce crystalliza- at 873 K for different time periods between 20 and 80 h. These
tion [251,255]. Amorphous TiO2 can be converted to anatase phase results were due to the formation of the surface-phase junction
due to pore collapse and crystal growth of the TiO2 powder during which was found to promote electron transfer from the conduction
calcination. With increasing temperatures, calcinations may result band of rutile to the trapping sites of anatase. Cybula et al. [234]
in phase transformation, reduced surface area, grain growth and synthesized titania nanoparticles using TiO2 P25 and found the
rate of CO2 photoreduction to methane was much higher on
O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42 35

temperature increase from 353 to 393 K and drying times from 5 to and graphitic carbon nitride [270,271]. In summary, these findings
20 h. A marked decrease in photoreduction efficiency was observed clearly indicate that the choice of catalyst precursors must be
on further temperature increase to 433 K and drying time of 35 h. carefully considered during catalyst synthesis with systematically
Asi et al. [75] synthesized AgBr doped TiO2 that exhibited good designed control experiment [226,266,272] or carbon source
crystallinity and optimal hydrocarbon production rate for the verification by GC–MS, NMR or LC–MS with isotope labelled 13CO2
photoreduction of CO2 under visible light at calcination tempera- in place.
ture of 773 K and doping concentration of 23.2 wt%. They found
that increasing the calcination temperature to 973 K resulted in the 5.2. Vapor deposition
aggregation of the doped nanoparticles which explained the
decreased in hydrocarbon production rate. Chemical vapor deposition (CVD) has been widely used in
surface coating in CO2 photoreduction and is prepared via the
5.1.2. Influence of organic contaminants vapor phase, while the formation of films in physical vapor
Organic impurities originating from chemical precursors used deposition (PVD) occurs when no chemical reaction takes place.
during catalyst synthesis have been reported to influence the Silija et al. [273] described CVD as a better technique when
activity and selectivity of CO2 photoreduction. These organic compared to PVD because better durability, adhesion and
compounds in the form of carbon and chlorine species which uniformity can be achieved with this technique compared to the
adsorb on the surface of TiO2 can serve as a carbon or chlorine latter. The need for aging, drying and reduction is also eliminated
source for the formation of hydrocarbons or undesired products with CVD. Extensive studies conducted by Choy [245] detailed
such as CH3Cl. By using a combination of in situ diffuse several variants of CVD and noted the complexity of chemical
reflectance infrared Fourier transform spectroscopy (DRIFTS) process as a major difference between CVD and physical vapor
with isotope labelled 13CO2, Yang et al. [264] demonstrated that deposition (PVD) due to the chemical precursors and reactions
carbon residues present on the catalyst surface were involved in used. The kind of precursor and substrate used with the desired
the photocatalytic reduction of CO2 to CO. It was observed that properties serves as a determining factor for the type of variant
prolonged exposure of the catalyst surface to UV irradiation and used. The thermal and chemical stability of the support and
H2O vapor was more effective for removal of these carbon operating conditions of the volatile precursors i.e., temperature
residues which originated from Ti alkoxides and polyethylene required for crystallization must be carefully considered. Thin films
glycol (PEG) than thermal treatment in air. Since adsorbed 12CO are usually formed in CVD when the surface of the substrate is
species were observed as the main product compared to 13CO exposed to volatile precursor(s) in inert atmosphere under
over Cu(I)/TiO2 in the absence of 12CO2, it was concluded that 12C controlled temperature and pressure. Nishida et al. [274]
originating from carbon residues was the predominant carbon demonstrated the use of plasma enhanced CVD for the preparation
source. Isotopic labelling results of Ag, Au and Pd–TiO2 samples of thin films of TiO2 while Galindo et al. [246] reported the use of
tested for CO2 reduction also confirmed that CH4 was formed pulsed injected metal organic CVD technique toward the deposi-
from organic impurities rather than 13CO2 [265]. The formation tion of multilayer thin films.
of chloromethane (CH3Cl) as a result of CO2 photoreduction with Wang et al. [213] synthesized platinized TiO2 films via the
Cu/I–TiO2 synthesized from a chlorinated precursor (CuCl2) was aerosol chemical vapor deposition (ACVD) technique. The synthe-
observed [181]. CH3Cl was formed from the reaction of methyl sized films which have unique one-dimensional structure of TiO2
radical (CH3) with chlorine radical (Cl.). Adsorbed carbon single crystals coated with Pt nanoparticles were reported to
species were reported to be intermediates for CH4 and CH3Cl exhibit high photoefficiency for CO2 reduction with water vapor
formation. following 4 h of UV irradiation. Maximum CH4 yield of 1361 mmol/
Several authors have employed different spectroscopic tech- gcatal h1 was exhibited by the Pt film with deposition time of 20 s.
niques such as DRIFTS, GC–MS, NMR or LC–MS with isotope Overall, high surface area, single crystallinity of the one
labelled 13CO2 as the reactant for verifying the carbon source of dimensional TiO2 films and efficient hole separation were the
final products generated from photocatalytic CO2 reduction. For main reasons described by the authors for the enhanced photo-
example, Yui et al. [266] observed that CO2 and CO32 were the activity of the films compared to TiO2. Asi et al. [75] demonstrated
main carbon sources of CH4 produced using Pd–TiO2 treated by the visible light reduction of CO2 to fuels using AgBr/CNT
calcination and washing procedures. Signal of m/z = 17 attributed nanocomposites. Multi-walled carbon nanotubes (CNT) were
to 13CH4 detected by the GC–MS when 13CO2 was used as a reactant synthesized by CVD while AgBr was introduced to the CNT
clearly demonstrated that CH4 formation was from CO2 photore- framework via deposition–precipitation method in the presence of
duction and not from residual carbon species. To verify the source cation surfactant. Higher product yields were obtained over AgBr/
of evolved CO and O2 from CO2 photoreduction with H2O, CNT nanocomposites compared to AgBr crystals. The product yield
Teramura et al. performed labelling experiments with 13CO2 and also increased with increasing nanotube length due to efficient
H218O as reactants using GC–MS [267]. After photo irradiation, charge transport demonstrated by longer nanotubes which was
peaks of 16O18O and 13CO were detected. The effect of several confirmed by electrochemical impedance spectroscopy measure-
solvents on CO2 photoreduction with Q-TiO2/SiO2 films was also ments.
studied by using 13C NMR and 13CO2 to identify the carbon source
for CO and C1–C2 oxygenates [268]. Labelling experiments 6. Photoreactor design and configuration
confirmed formate and CO were produced from CO2 and not from
the solvents (CCl4, 2-propanol and acetonitrile). Fu et al. used The configuration of catalyst particles in a photoreactor
isotopic 13CO2 for the photocatalytic reduction of CO2 over system is also another important factor that can influence the
titanium metal organic framework (MOF) catalysts where overall photocatalytic efficiency [27,28]. Photoreactors are vessels
obtained products were identified by 13C NMR spectroscopy where reaction products are generated from the contact between
[269]. Isotopic 13CO2 reaction confirmed that the reduced product, photocatalysts, reactants and photons. The two key parameters
HCOO originated from CO2 rather than residual carbon species. which determine the types of photoreactors utilized in CO2
Ohno et al. also demonstrated that that CO2 was the carbon source photoreduction are the phases involved (i.e., multiphase (gas–
for CH3OH evolution with 1H NMR spectroscopy over Au–TiO2 solid, liquid–solid, gas–liquid–solid etc.) and the mode of
(brookite) nanorods and hybrid photocatalyst composed of WO3 operation (i.e., batch, semi-batch or continuous) [158].
36 O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42

Photocatalysts can be generally tested in either suspended or


immobilized forms in reactor configurations, as shown in Fig. 10.
An ideal photoreactor must have uniform light distribution
throughout the entire system in order to achieve optimum
results. Currently, comparative analysis of product yield and
reactor configurations in CO2 reduction is primarily reported in
terms of quantum efficiency. The advantages and disadvantages
of different types of photoreactor systems currently used in CO2
photoreduction are summarized in Table 4 [275–279].

6.1. Fluidized and slurry reactor designs

When powders are dispersed/suspended in a liquid medium;


the quantum efficiency of the catalyst, absorption properties of
both reactants and non-reactants in solution and surface light
intensity determines the rate of reaction [28]. Some of the key
advantages of slurry system are that there is entire external surface
illumination during reaction time if the particle size of the catalyst
is small with phase segregation not occurring if the solution is
homogeneously mixed [27]. Although slurry system designs offer
high catalyst loading and ease of construction; separation of
catalyst particles from the reaction mixture is a major drawback
[221,279]. The size of the catalyst crystals or aggregates (0.05 mm
to a few mm) will determine the nature of separation process
required which could be expensive and time consuming [158,279].
However, the penetration depth of UV light into the reaction
medium can also be limited by the strong light absorption of
organic species and catalyst particles [28,219]. A large proportion
of catalyst surface area might be inactive due to low photon energy
received, as most of the light irradiation is lost due to absorption by
liquid when light approaches the catalyst through the bulk liquid
phase [280,281]. Light distribution can be better controlled via
external or internal illumination. Hydrocarbon formation (CO,
Fig. 10. Schematic of (A) slurry reactor design with top illumination, (B) optical CH3CHO and CH3CHO2) was observed over hybrid catalyst, TiO2:
fiber reactor design with side illumination and (C) internally illuminated reactor Rh-LHCII tested in a stirred batch reactor under visible light
with top illumination.
irradiation [272]. Hybrid catalyst, TiO2:Rh-LHCII was formulated
by attaching light-harvesting complexes (LHCII) extracted from
spinach to the surface of Rh-doped TiO2 (TiO2:Rh) in order to

Table 4
Advantages and disadvantages of photoreactor systems.

Reactor design Advantages Disadvantages References


Fluidized and slurry (I) Temperature gradients inside the beds can be reduced through (I) Filters (liquid phase) and scrubbers (gas) are needed [13]
reactor vigorous movements caused by the solid passing through the fluids
(multiphase)
(II) Heat and mass transfer rates increase considerable due to agitated (II) Flooding tends to reduce the effectiveness of the
movement of solid particles catalyst
(III) High catalyst loading (III) Difficulty of separating the catalyst from the reaction [219,220]
mixture
(IV) Low light utilization efficiency due to absorption and
scattering of the light by the reaction medium
V) Restricted processing capacities due to mass transport
limitations

Fixed bed reactor (I) High surface area (I) Temperature gradient between gas and solid surface is [275,280]
(II) Fast reaction time common
(III) The conversion rate per unit mass of the catalyst is high due to the
flow regime close to plug flow
(IV) Low operating costs due to low pressure drop

Variants of fixed bed designs


Monolith reactor (I) High surface to volume ratio and low pressure drop with high flow rate (I) Low light efficiency due to opacity of channels of the [276,277]
can be achieved monolith
(II) Configuration can be easily modified

Optical fiber reactor (I) High surface area and light utilization efficiency (I) Maximum use of the reactor volume is not achieved [277,278]
(II) Efficient processing capacities of the catalyst (II) Heat build-up of fibers can lead to rapid catalyst
deactivation
O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42 37

enhance the light absorption and increase yields. The quantum reduction of CO2 for 1 wt% Pd/0.01 wt% Rh–TiO2 in a slurry batch
efficiency of 0.0411% was achieved using this configuration. annular reactor system and internally illuminated photoreactor
Slurry reactor design with two separate components for oxygen system was evaluated [155]. The quantum efficiency of the
and hydrocarbon evolution separated by a Nafion membrane was internally illuminated monolith reactor (0.049%) was near one
tested by Lee at al. [107] under AM 1.5G irradiation. WO3 was order of magnitude higher than the slurry batch annular reactor
chosen as the oxidation catalyst while Pt/CuGaAlO4 and Pt/SrTiO3: (0.002%). The performance of TiO2 for CO production was evaluated
Rh as reduction catalyst. The dual photocatalyst system containing in cell type and multichannel monolith reactor [226]. The
both reduction and oxidation catalysts demonstrated higher performance comparison between the gas phase photoreactors
quantum efficiency of 0.0051% compared to the single photo- revealed 8 fold higher yield of CO in the monolith compared to cell
catalyst system with Pt/CuGaAlO4 as reduction catalyst only type reactor. Quantum efficiency in microchannel monolith reactor
(0.0019%). Photocatalytic reduction of CO2 with monoethanol- was 0.0301% compared to 0.0028% in the cell type reactor.
amine (MEA) as reductants to form CH4 using Ti-MCM-41 meso- Improved quantum efficiency was reported to be due to higher
porous catalysts was studied and compared to other reductants illuminated surface area, higher photon energy consumption and
such as NaOH and H2O in slurry reactor designs [184]. Photore- better utilization of reactor volume.
duction efficiency of Ti-MCM-41(50) sample 8 h of UV illumination
using H2O, NaOH and MEA as reductants were 0.83, 0.29 and 9.18%, 7. Conclusions and future perspectives
respectively. Copper or cobalt incorporated TiO2 supported ZSM-
5 catalysts were tested in a slurry photoreactor where 0.1 M The utilization of CO2 as a direct feedstock for photocatalytic
NaHCO3 was the reductant [282]. When Cu–TiO2/ZSM-5 was used conversion into fuels over different variants of pure and modified
as the catalyst, the quantum efficiency of CH3OH reached 10.11% TiO2 synthesized by various routes and tested in various photo-
after 12 h reaction time. High absorption ability of ZSM-5 was reactor designs has been highlighted in this review. Application of
reported to influence the photoconversion rates. TiO2 induced photocatalysis for the challenging conversion of CO2
remains a promising pathway as it can be activated by solar energy
6.2. Fixed bed reactor designs at relatively mild conditions to form valuable products. Although
recent progress focused on the use of pure and modified
The drawback of catalyst separation can be avoided by fixed bed photoactivated TiO2 materials has induced fuel generation from
reactor designs where catalysts are immobilized onto fixed CO2 reduction with H2O; expected improvement required for
supports e.g., plate, beads, fibers, monolith etc. In these systems, scalable fuel production has not been achieved. To this end, design
photocatalysts are coated on a support matrix placed inside the and synthesis of novel TiO2 photocatalysts with higher stability,
reactor around the light source or directly on the photoreactor selectivity and efficiency requires improvements in synthetic
wall. Light distribution becomes a limiting factor in this system procedures offering improved control over physicochemical
which is influenced by the geometry of the light source and spatial properties. Analytical techniques such as in-situ surface and bulk
distance between this light source and photocatalysts [197]. spectroscopies must be employed to provide valuable insight into
Overall, gas phase systems offer more flexibility compared to the fundamental steps occurring in CO2 photocatalytic reduction, rate
slurry systems if the design considers the spatial relationship limiting steps, formation and stability of surface reaction
between the reactor and the light source when choosing the intermediates as well as adsorption and desorption of both
support of choice. Several researchers have designed photoreactors reactant and product species. Besides the materials science aspect
using optical fibers as supports [217,278,283]. The conventional of CO2 photocatalytic reduction, the engineering challenge of
optical fiber reactor (OFR) has been modified by using fibers with optimal CO2 photoreactor design needs a step change transforma-
different cores and coatings [218,284], increasing their diameter to tion to reach its crucial role in the overall process performance.
create ease of handling and the use of cooling systems [285] to From this review, it can be concluded that CO2 photocatalysis is still
eliminate the limitation of heat build-up. Wu et al. [154,286] have not feasible due to the absence of scalable reactor designs able to
conducted CO2 photoreduction studies using the optical fiber simultaneously introduce reactants, light and efficient visible light
reactor system. They coated optical fibers with Cu/TiO2 and responsive catalysts to effect production of specific fuels in
Ag/TiO2 catalysts in the gaseous phase, respectively. A maximum significant quantities. Since the overall process efficiency is largely
yield of 0.91 mmol/gcatal h was observed using the loading ratio of dependent on two factors—the reactor configuration and physico-
0.5 wt% Cu–Fe/TiO2 for methane production. Maximum methanol chemical properties of the catalyst, it is desirable to scale up this
yield of 0.45 mmol/gcatal h was observed by Wu et al. [286] using system based on the design and development of these parameters
1.2 wt% Cu/TiO2 under light intensity of 16 W/cm2 while methanol incorporating maximal light efficiency and mass transfer. Perfor-
yield of 4.12 mmol/gcatal h was observed by Wu et al. [154] using mance of current photoreactor designs is primarily reported and
1 wt% Ag/TiO2 under light intensity of 10 W/cm2. In particular, evaluated in terms of quantum efficiency without consideration of
previous CO2 photoreduction studies conducted by Nguyen and light transport to active site which is critical for scale-up and
Wu [219] using optical fibers coated with Cu–Fe/TiO2 catalysts in quantification of energy losses due to light absorption by reaction
the gaseous phase have demonstrated that the number of optical media and reactor components. A deep understanding of
fibers can determine the rate of ethylene production and engineering aspects of CO2 reduction is still required for the
selectively increase or decrease the quantum yield. A maximum development of highly efficient photoreactor designs. In order to
yield of 0.91 mmol/gcatal h was observed using the loading ratio of achieve high conversion efficiency, photoreactor designs must take
0.5 wt% Cu–Fe/TiO2 for CH4 production. CO2 photocatalytic activity into account the material of construction, its thickness, mass of
of NiO/InTaO4 catalysts dispersed in aqueous solution of NaOH catalyst, reactor geometry (length, volume etc.), flow rate and the
(slurry designs) and immobilized in a fixed bed reactor design relationship between the reactor and irradiation source. The
containing 216 optical fibers was evaluated by Wang et al. [193]. modelling of the effect of reactor designs and operation parameters
The quantum efficiency for methanol production was 14 times on CO2 reduction is also required to extrapolate results for the
higher in optical-fiber reactor (0.053%) than that of the aqueous- design of pilot scale systems. Furthermore, this work can be
phase reactor (0.0045%). The higher quantum efficiency was extended to include the use of flue gas generated from power
attributed to improved light efficiency by the films coated on plants as a feedstock for CO2 reduction. Different compositions of
optical fibers. The comparison between the photocatalytic flue gas streams can be used directly or indirectly in order to
38 O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42

ascertain the effect of impurities and the concentration of CO2 [26] M. Sahu, P. Biswas, Single-step processing of copper-doped titania
required to achieve maximum conversion rates. Results from using nanomaterials in a flame aerosol reactor, Nanoscale Res. Lett. 6 (1) (2011) 1–
14.
concentrated CO2 gas streams derived from the flue gas can also be [27] M. Bideau, et al., On the immobilization of titanium dioxide in the
tested and compared to pure flue gas streams to determine the photocatalytic oxidation of spent waters, J. Photochem. Photobiol. A 91 (2)
most suitable option for scalable fuel production. (1995) 137–144.
[28] A.K. Ray, A.A. Beenackers, Development of a new photocatalytic reactor for
water purification, Catal. Today 40 (1) (1998) 73–83.
Acknowledgements [29] Z. Jiang, et al., Turning carbon dioxide into fuel, Philos. Trans. R. Soc. A 368
(1923) (2010) 3343–3364.
[30] U. Akpan, B. Hameed, Parameters affecting the photocatalytic degradation of
The authors thank the financial support provided by the School dyes using TiO2-based photocatalysts: a review, J. Hazard. Mater. 170 (2)
of Engineering and Physical Sciences and the Centre for Innovation (2009) 520–529.
in Carbon Capture and Storage (EPSRC grant number EP/K021796/ [31] K. Ko9 cí, L. Obalová, Z. Lacný, Photocatalytic reduction of CO2 over TiO2 based
catalysts, Chem. Papers 62 (1) (2008) 1–9.
1) at Heriot-Watt University.
[32] A.L. Linsebigler, G. Lu, J.T. Yates Jr, Photocatalysis on TiO2 surfaces: principles,
mechanisms, and selected results, Chem. Rev. 95 (3) (1995) 735–758.
[33] V.P. Indrakanti, J.D. Kubicki, H.H. Schobert, Photoinduced activation of CO2 on
References Ti-based heterogeneous catalysts: current state, chemical physics-based
insights and outlook, Energy Environ. Sci. 2 (7) (2009) 745–758.
[1] IPCC. Special Report on Renewable Energy Sources and Climate Change [34] Marinkovic, et al., Modern Aspects of Electrochemistry, vol. 42, Springer, New
Mitigation. [cited 2014 November 19]. http://www.ipcc.ch/report/srren/ York, 2008.
(2011). [35] S.N. Habisreutinger, L. Schmidt-Mende, J.K. Stolarczyk, Photocatalytic
[2] C. Global, The Global Status Of CCS, Institute (GCCSI), Canberra, Australia, reduction of CO2 on TiO2 and other semiconductors, Angew. Chem. Int. Ed. 52
2012. (29) (2013) 7372–7408.
[3] P. Moriarty, D. Honnery, Mitigating greenhouse: limited time, limited [36] P.D. Tran, et al., Recent advances in hybrid photocatalysts for solar fuel
options, Energy Policy 364 (2008) 1251–1256. production, Energy Environ. Sci. 5 (3) (2012) 5902–5918.
[4] J.O.M. Bockris, Would methanol formed from CO2 from the atmosphere give [37] K. Wang, et al., Sulfur-doped g-C3N4 with enhanced photocatalytic CO2-
the advantage of hydrogen at lesser cost? Int. J. Hydrogen Energy 35 (11) reduction performance, Appl. Catal. B 176–177 (2015) 44–52.
(2010) 5165–5172. [38] B. Wang, H. Yang, T. Xian, L. Di, J.R.S. Li, X.X. Wang, Synthesis of spherical
[5] C. Song, Global challenges and strategies for control, conversion and Bi2WO6 nanoparticles by a hydrothermal route and their photocatalytic
utilization of CO2 for sustainable development involving energy, catalysis, properties, J. Nanomater. (2015) 1–7.
adsorption and chemical processing, Catal. Today 115 (1) (2006) 2–32. [39] A. Mills, S. Le Hunte, An overview of semiconductor photocatalysis, J.
[6] J.A. Turner, A realizable renewable energy future, Science 285 (5428) (1999) Photochem. Photobiol. A 108 (1) (1997) 1–35.
687–689. [40] D. Beydoun, et al., Role of nanoparticles in photocatalysis, J. Nanoparticle Res.
[7] M. Grimston, et al., The European and global potential of carbon dioxide 1 (4) (1999) 439–458.
sequestration in tackling climate change, Clim. Policy 1 (2) (2001) 155–171. [41] S.M. Gupta, M. Tripathi, A review of TiO2 nanoparticles, Chin. Sci. Bull. 56 (16)
[8] M. Olivares-Marín, M.M. Maroto-Valer, Development of adsorbents for CO2 (2011) 1639–1657.
capture from waste materials: a review, Greenhouse Gases Sci. Technol. 2 (1) [42] M.A. Fox, M.T. Dulay, Heterogeneous photocatalysis, Chem. Rev. 93 (1) (1993)
(2012) 20–35. 341–357.
[9] S. Bachu, et al., CO2 storage capacity estimation: methodology and gaps, Int. J. [43] D.W. Bahnemann, C. Kormann, M.R. Hoffmann, Preparation and
Greenhouse Gas Control 1 (4) (2007) 430–443. characterization of quantum size zinc oxide: a detailed spectroscopic study, J.
[10] S. Bachu, CO2 storage in geological media: role, means, status and barriers to Phys. Chem. 91 (14) (1987) 3789–3798.
deployment, Prog. Energy Combust. Sci. 34 (2) (2008) 254–273. [44] V. Jeyalakshmi, et al., Photocatalytic reduction of carbon dioxide by water: a
[11] M. Aresta, A. Dibenedetto, Utilisation of CO2 as a chemical feedstock: step towards sustainable fuels and chemicals, Materials Science Forum, Trans
opportunities and challenges, Dalton Trans. 28 (2007) 2975–2992. Tech Publication, 2013.
[12] K. Koci, L. Obalova, Z. Lacny, Photocatalytic reduction of CO2 over TiO2 based [45] U. Diebold, The surface science of titanium dioxide, Surf. Sci. Rep. 48 (5–8)
catalysts, Chem. Papers 62 (1) (2008) 1–9. (2003) 53–229.
[13] P. Usubharatana, et al., Photocatalytic process for CO2 emission reduction [46] L. Chen, et al., Photoreduction of CO2 by TiO2 nanocomposites synthesized
from industrial flue gas streams, Ind. Eng. Chem. Res. 45 (8) (2006) 2558– through reactive direct current magnetron sputter deposition, Thin Solid
2568. Films 517 (19) (2009) 5641–5645.
[14] M. Mikkelsen, M. Jorgensen, F.C. Krebs, The teraton challenge. A review of [47] Y. Kohno, et al., Photoreduction of CO2 with H2 over ZrO2. A study on
fixation and transformation of carbon dioxide, Energy Environ. Sci. 3 (1) interaction of hydrogen with photoexcited CO2, Phys. Chem. Chem. Phys. 2
(2010) 43–81. (11) (2000) 2635–2639.
[15] H. Yano, et al., Efficient electrochemical conversion of CO2 to CO, C2H4 and [48] T.V. Nguyen, J.C.S. Wu, Photoreduction of CO2 to fuels under sunlight using
CH4 at a three-phase interface on a Cu net electrode in acidic solution, J. optical-fiber reactor, Sol. Energy Mater. Sol. Cells 92 (8) (2008) 864–872.
Electroanal. Chem. 519 (1–2) (2002) 93–100. [49] G. Guan, T. Kida, A. Yoshida, Reduction of carbon dioxide with water under
[16] J. Yano, et al., Selective ethylene formation by pulse-mode electrochemical concentrated sunlight using photocatalyst combined with Fe-based catalyst,
reduction of carbon dioxide using copper and copper-oxide electrodes, J. Appl. Catal. B 41 (4) (2003) 387–396.
Solid State Electrochem. 11 (4) (2007) 554–557. [50] S. Qin, et al., Photocatalytic reduction of CO2 in methanol to methyl formate
[17] Creutz, F. Fujita, Carbon dioxide as a feedstock, Carbon Management: over CuO–TiO2 composite catalysts, J. Colloid Interface Sci. 356 (1) (2011)
Implications for R&D in the Chemical Sciences and Technology: A Workshop 257–261.
Report to the Chemical Sciences Roundtable, National Academies Press, [51] L. Jia, J. Li, W. Fang, Enhanced visible-light active C and Fe co-doped LaCoO3
Washington, D.C., US, 2001. for reduction of carbon dioxide, Catal. Commun. 11 (2) (2009) 87–90.
[18] M.M. Maroto-Valer, Developments And Innovation In Carbon Dioxide (CO2) [52] K. Teramura, et al., Effect of H2 gas as a reductant on photoreduction of CO2
Capture And Storage Technology: Carbon Dioxide (CO2) Storage And over a Ga2O3 photocatalyst, Chem. Phys. Lett. 467 (1–3) (2008) 191–194.
Utilization, Taylor & Francis, 2010, 2015. [53] K. Teramura, et al., Photocatalytic reduction of CO2 using H2 as reductant over
[19] O. Carp, C.L. Huisman, A. Reller, Photoinduced reactivity of titanium dioxide, ATaO3 photocatalysts (A = Li, Na, K), App. Catal. B 96 (3–4) (2010) 565–568.
Prog. Solid State Chem. 32 (1–2) (2004) 33–177. [54] J.C. Colmenares, et al., Nanostructured photocatalysts and their applications
[20] Slamet, et al., Photocatalytic reduction of CO2 on copper-doped titania in the photocatalytic transformation of lignocellulosic biomass: an overview,
catalysts prepared by improved-impregnation method, Catal. Commun. 6 (5) Materials 2 (4) (2009) 2228–2258.
(2005) 313–319. [55] M. Bellardita, et al., Photocatalytic behaviour of metal-loaded TiO2 aqueous
[21] N. Murakami, et al., Photocatalytic reduction of carbon dioxide over shape- dispersions and films, Chem. Phys. 339 (1–3) (2007) 94–103.
controlled titanium(IV) oxide nanoparticles with co-catalyst loading, Curr. [56] S. Malato, et al., Decontamination and disinfection of water by solar
Org. Chem. 17 (21) (2013) 2449–2453. photocatalysis: recent overview and trends, Catal. Today 147 (1) (2009) 1–59.
[22] O. Ishitani, et al., Photocatalytic reduction of carbon dioxide to methane and [57] L. Yuan, Y.-J. Xu, Photocatalytic conversion of CO2 into value-added and
acetic acid by an aqueous suspension of metal-deposited TiO2, J. Photochem. renewable fuels, Appl. Surf. Sci. 342 (2015) 154–167.
Photobiol. A 72 (3) (1993) 269–271. [58] M. Marszewski, S. Cao, J. Yu, M. Jaroniec, Semiconductor-based photocatalytic
[23] F. Solymosi, I. Tombacz, Photocatalytic reaction of H2O + CO2 over pure and CO2 conversion, Mater. Horizons 2 (3) (2015) 261–278.
doped Rh/TiO2, Catal. Lett. 27 (1–2) (1994) 61–65. [59] X. Li, et al., Design and fabrication of semiconductor photocatalyst for
[24] R. Asahi, et al., Visible-light photocatalysis in nitrogen-doped titanium photocatalytic reduction of CO2 to solar fuel, Sci. China Mater. 57 (1) (2014)
oxides, Science 293 (5528) (2001) 269–271. 70–100.
[25] Q. Zhang, et al., Visible light responsive iodine-doped TiO2 for photocatalytic [60] V.P. Indrakanti, Kubicki, H.H. Schobert, Photoinduced activation of CO2 on Ti-
reduction of CO2 to fuels, Appl. Catal. A 400 (1–2) (2011) 195–202. based heterogeneous catalysts: current state, chemical physics-based
insights and outlook, Energy Environ. Sci. 2 (7) (2009) 745–758.
O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42 39

[61] G. Li Puma, et al., Preparation of titanium dioxide photocatalyst loaded onto [93] H.C. Yang, et al., Mesoporous TiO2/SBA-15, and Cu/TiO2/SBA-15 composite
activated carbon support using chemical vapor deposition: a review paper, J. photocatalysts for photoreduction of CO2 to methanol, Catal. Lett. 131 (3–4)
Hazard. Mater. 157 (2–3) (2008) 209–219. (2009) 381–387.
[62] P. Bouras, E. Stathatos, P. Lianos, Pure versus metal-ion-doped [94] Y. Li, et al., Photocatalytic reduction of CO2 with H2O on mesoporous silica
nanocrystalline titania for photocatalysis, Appl. Catal. B 73 (1–2) (2007) 51– supported Cu/TiO2 catalysts, Appl. Catal. B 100 (1–2) (2010) 386–392.
59. [95] K. Kocí, et al., Comparison of the pure TiO2 and kaolinite/TiO2 composite as
[63] W.A. Zeltner, D.T. Tompkins, L. Harriman, Shedding light on photocatalysis. catalyst for CO2 photocatalytic reduction, Catal. Today 161 (1) (2011) 105–
Discussion, ASHRAE Trans. (2005) 523–534. 109.
[64] S. Watson, et al., Preparation of nanosized crystalline TiO2 particles at low [96] J. Wei, et al., Characterization of Y/TiO2 nanoparticles and their reactivity for
temperature for photocatalysis, J. Nanoparticle Res. 6 (2) (2004) 193–207. CO2 photoreduction, Appl. Mech. Mater. 55 (2011) 1506–1510.
[65] A. Fujishima, X. Zhang, D.A. Tryk, TiO2 photocatalysis and related surface [97] C. Zhao, et al., Photocatalytic conversion of CO2 and H2O to fuels by
phenomena, Surf. Sci. Rep. 63 (12) (2008) 515–582. nanostructured Ce–TiO2/SBA-15 composites, Catal. Sci. Technol. 2 (12) (2012)
[66] M. Kitano, et al., Recent developments in titanium oxide-based 2558–2568.
photocatalysts, Appl. Catal. A 325 (1) (2007) 1–14. [98] L. Collado, et al., Enhancement of hydrocarbon production via artificial
[67] F. Han, et al., Tailored titanium dioxide photocatalysts for the degradation of photosynthesis due to synergetic effect of Ag supported on TiO2 and ZnO
organic dyes in wastewater treatment: a review, Appl. Catal. A 359 (1) (2009) semiconductors, Chem. Eng. J. 224 (2013) 128–135.
25–40. [99] B.S. Kwak, et al., Methane formation from photoreduction of CO2 with water
[68] O. Ola, M.M. Maroto-Valer, Transition metal oxide based TiO2 nanoparticles using TiO2 including Ni ingredient, Fuel 143 (2015) 570–576.
for visible light induced CO2 photoreduction, Appl. Catal. A 502 (2015) 114– [100] Y. Liu, et al., Photocatalytic reduction of CO2 with water vapor on surface La-
121. modified TiO2 nanoparticles with enhanced CH4 selectivity, Appl. Catal. B
[69] X. Chen, Titanium dioxide nanomaterials: synthesis, properties, 168–169 (2015) 125–131.
modifications, and applications, Cheminform 38 (41) (2007) . [101] C. Tang, et al., CeF3/TiO2 composite as a novel visible-light-driven
[70] O. Ozcan, et al., Dye sensitized artificial photosynthesis in the gas phase over photocatalyst based on upconversion emission and its application for
thin and thick TiO2 films under UV and visible light irradiation, Appl. Catal. B photocatalytic reduction of CO2, J. Lumin. 154 (2014) 305–309.
Environ. 71 (3–4) (2007) 291–297. [102] O. Ola, M. Mercedes Maroto-Valer, Role of catalyst carriers in CO2
[71] T.-V. Nguyen, J.C.S. Wu, C.-H. Chiou, Photoreduction of CO2 over ruthenium photoreduction over nanocrystalline nickel loaded TiO2-based
dye-sensitized TiO2-based catalysts under concentrated natural sunlight, photocatalysts, J. Catal. 309 (2014) 300–308.
Catal. Commun. 9 (10) (2008) 2073–2076. [103] S. Rani, N. Bao, S.C. Roy, Solar spectrum photocatalytic conversion of CO2 and
[72] K.R. Thampi, J. Kiwi, M. Graetzel, Methanation and photo-methanation of water vapor into hydrocarbons using TiO2 nanoparticle membranes, Appl.
carbon dioxide at room temperature and atmospheric pressure, Nature 327 Surf. Sci. 289 (2014) 203–208.
(6122) (1987) 506–508. [104] L. Matejova, et al., Preparation, characterization and photocatalytic
[73] G. Qin, et al., Photocatalytic reduction of carbon dioxide to formic acid, properties of cerium doped TiO2: on the effect of Ce loading on the
formaldehyde, and methanol using dye-sensitized TiO2 film, Appl. Catal. B photocatalytic reduction of carbon dioxide, Appl. Catal. B 152 (153) (2014)
129 (2013) 599–605. 172–183.
[74] C.J. Wang, et al., Visible light photoreduction of CO2 using CdSe/Pt/TiO2 [105] Y. Wang, et al., High efficiency photocatalytic conversion of CO2 with H2O
heterostructured catalysts, J. Phys. Chem. Lett. 1 (1) (2010) 48–53. over Pt/TiO2 nanoparticles, RSC Adv. 4 (84) (2014) 44442–44451.
[75] M. Abou Asi, et al., Photocatalytic reduction of CO2 to hydrocarbons using [106] P. Akhter, et al., New nanostructured silica incorporated with isolated Ti
AgBr/TiO2 nanocomposites under visible light, Catal. Today 175 (1) (2011) material for the photocatalytic conversion of CO2 to fuels, Nanoscale Res. Lett.
256–263. 9 (1) (2014) 1–8.
[76] C. Wang, et al., Size-dependent photocatalytic reduction of CO2 with PbS [107] W.-H. Lee, et al., A novel twin reactor for CO2 photoreduction to mimic
quantum dot sensitized TiO2 heterostructured photocatalysts, J. Mater. artificial photosynthesis, Appl. Catal. B 132 (2013) 445–451.
Chem. 21 (35) (2011) 13452–13457. [108] Y.-H. Cheng, et al., Photo-enhanced hydrogenation of CO2 to mimic
[77] X. Li, et al., Adsorption of CO2 on heterostructure CdS (Bi2S3)/TiO2 nanotube photosynthesis by CO co-feed in a novel twin reactor, Appl. Energy 147 (2015)
photocatalysts and their photocatalytic activities in the reduction of CO2 to 318–324.
methanol under visible light irradiation, Chem. Eng. J. 180 (2012) 151–158. [109] F. Saladin, L. Forss, I. Kamber, Photosynthesis of CH4 at a TiO2 surface from
[78] Y. Wang, et al., Ordered mesoporous CeO2–TiO2 composites: highly efficient gaseous H2O and CO2, J. Chem. Soc. Chem. Commun. 5 (1995) 533–534.
photocatalysts for the reduction of CO2 with H2O under simulated solar [110] K. Iizuka, et al., Photocatalytic reduction of carbon dioxide over Ag cocatalyst-
irradiation, Appl. Catal. B 130 (2013) 277–284. loaded ALa4Ti4O15 (A = Ca, Sr, and Ba) using water as a reducing reagent, J. Am.
[79] A.A. Beigi, S. Fatemi, Z. Salehi, Synthesis of nanocomposite CdS/TiO2 and Chem. Soc. 133 (51) (2011) 20863–20868.
investigation of its photocatalytic activity for CO2 reduction to CO and CH4 [111] K. Ogura, et al., Visible-light-assisted decomposition of H2O and
under visible light irradiation, J. CO2 Util. 7 (2014) 23–29. photomethanation of CO2 over CeO2–TiO2 catalyst, J. Photochem. Photobiol. A
[80] K. Ko9cí, et al., Sol–gel derived Pd supported TiO2–ZrO2 and TiO2 66 (1) (1992) 91–97.
photocatalysts; their examination in photocatalytic reduction of carbon [112] Q. Li, et al., Photocatalytic reduction of CO2 on MgO/TiO2 nanotube films,
dioxide, Catal. Today 230 (2014) 20–26. Appl. Surf. Sci. 314 (2014) 458–463.
[81] G. Marcì, E.I. García-López, L. Palmisano, Photocatalytic CO2 reduction in gas– [113] M. Tahir, N.S. Amin, Indium-doped TiO2 nanoparticles for photocatalytic CO2
solid regime in the presence of H2O by using GaP/TiO2 composite as reduction with H2O vapors to CH4, Appl. Catal. B 162 (2015) 98–109.
photocatalyst under simulated solar light, Catal. Commun. 53 (2014) 38–41. [114] M. Hussain, et al., Nanostructured TiO2/KIT-6 catalysts for improved
[82] N. Sasirekha, S.J.S. Basha, K. Shanthi, Photocatalytic performance of Ru doped photocatalytic reduction of CO2 to tunable energy products, Appl. Catal. B
anatase mounted on silica for reduction of carbon dioxide, Appl. Catal. B 62 170–171 (2015) 53–65.
(1–2) (2006) 169–180. [115] J. Fan, et al., Synergistic effect of N and Ni2+ on nanotitania in photocatalytic
[83] S.S. Tan, L. Zou, E. Hu, Photosynthesis of hydrogen and methane as key reduction of CO2, J. Environ. Eng. 137 (3) (2010) 171–176.
components for clean energy system, Sci. Technol. Adv. Mater. 8 (1) (2007) [116] O.K. Varghese, et al., High-rate solar photocatalytic conversion of CO2 and
89–92. water vapor to hydrocarbon fuels, Nano Lett. 9 (2) (2009) 731–737.
[84] J.C.S. Wu, et al., Application of optical-fiber photoreactor for CO2 [117] Z. Zhao, et al., Effect of heating temperature on photocatalytic reduction of
photocatalytic reduction, Top. Catal. 47 (3–4) (2008) 131–136. CO2 by N–TiO2 nanotube catalyst, Catal. Commun. 21 (2012) 32–37.
[85] J. Wu, Photocatalytic reduction of greenhouse gas CO2 to fuel, Catal. Surv. Asia [118] L.M. Xue, et al., Preparation of C doped TiO2 photocatalysts and their
13 (1) (2009) 30–40. photocatalytic reduction of carbon dioxide, Adv. Mater. Res. 183 (2011) 1842–
[86] H.W.N. Slamet, et al., Effect of copper species in a photocatalytic synthesis of 1846.
methanol from carbon dioxide over copper-doped titania catalysts, World [119] B. Michalkiewicz, et al., Reduction of CO2 by adsorption and reaction on
Appl. Sci. J. 6 (1) (2009) 112–122. surface of TiO2-nitrogen modified photocatalyst, J. CO2 Util. 5 (2014) 47–52.
[87] I.H. Tseng, W.C. Chang, J.C.S. Wu, Photoreduction of CO2 using sol–gel derived [120] S. Zhou, et al., Facile in situ synthesis of graphitic carbon nitride (g-C3N4)–N–
titania and titania-supported copper catalysts, Appl. Catal. B 37 (1) (2002) TiO2 heterojunction as an efficient photocatalyst for the selective
37–48. photoreduction of CO2 to CO, Appl. Catal. B 158–159 (2014) 20–29.
[88] I.H. Tseng, J.C.S. Wu, H.Y. Chou, Effects of sol–gel procedures on the [121] T. Phongamwong, M. Chareonpanich, J. Limtrakul, Role of chlorophyll in
photocatalysis of Cu/TiO2 in CO2 photoreduction, J. Catal. 221 (2) (2004) 432– spirulina on photocatalytic activity of CO2 reduction under visible light over
440. modified N-doped TiO2 photocatalysts, Appl. Catal. B 168–169 (2015) 114–
[89] S.S. Tan, L. Zou, E. Hu, Photocatalytic reduction of carbon dioxide into gaseous 124.
hydrocarbon using TiO2 pellets, Catal. Today 115 (1–4) (2006) 269–273. [122] M. Ni, et al., A review and recent developments in photocatalytic water-
[90] S. Krej9cíková, et al., Preparation and characterization of Ag-doped crystalline splitting using TiO2 for hydrogen production, Renew. Sustain. Energy Rev. 11
titania for photocatalysis applications, Appl. Catal. B 111 (2012) 119–125. (3) (2007) 401–425.
[91] Q.-H. Zhang, et al., Photocatalytic reduction of CO2 with H2O on Pt-loaded [123] K. Kalyanasundaram, M. Graetzel, Artificial photosynthesis: biomimetic
TiO2 catalyst, Catal. Today 148 (3–4) (2009) 335–340. approaches to solar energy conversion and storage, Curr. Opin. Biotechnol. 21
[92] T.-f. Xie, et al., Application of surface photovoltage technique in (3) (2010) 298–310.
photocatalysis studies on modified TiO2 photo-catalysts for photo-reduction [124] Y. Cho, et al., Visible light-induced degradation of carbon tetrachloride on
of CO2, Mater. Chem. Phys. 70 (1) (2001) 103–106. dye-sensitized TiO2, Environ. Sci. Technol. 35 (5) (2001) 966–970.
40 O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42

[125] H. Ross, J. Bendig, S. Hecht, Sensitized photocatalytical oxidation of [159] E. Liu, et al., Photoconversion of CO2 to methanol over plasmonic Ag/TiO2
terbutylazine, Sol. Energy Mater. Sol. Cells 33 (4) (1994) 475–481. nano-wire films enhanced by overlapped visible-light-harvesting
[126] G. Mele, et al., Photocatalytic degradation of 4-nitrophenol in aqueous nanostructures, Ceram. Int. 41 (1) (2015) 1049–1057.
suspension by using polycrystalline TiO2 samples impregnated with Cu(II)- [160] B. Tahir, M. Tahir, N.S. Amin, Gold–indium modified TiO2 nanocatalysts for
phthalocyanine, Appl. Catal. B 38 (4) (2002) 309–319. photocatalytic CO2 reduction with H2 as reductant in a monolith
[127] M. Grätzel, Perspectives for dye-sensitized nanocrystalline solar cells, Prog. photoreactor, Appl. Surf. Sci. 338 (2015) 1–14.
Photovoltaics Res. Appl. 8 (1) (2000) 171–185. [161] S. Linic, P. Christopher, D.B. Ingram, Plasmonic-metal nanostructures for
[128] S.G. Kumar, L.G. Devi, Review on modified TiO2 photocatalysis under UV/ efficient conversion of solar to chemical energy, Nat. Mater. 10 (12) (2011)
visible light: selected results and related mechanisms on interfacial charge 911–921.
carrier transfer dynamics, J. Phys. Chem. A 115 (46) (2011) 13211–13241. [162] M. Ovcharov, V. Shvalagin, V. Granchak, Photocatalytic reduction of carbon
[129] Y.-J. Yuan, et al., A copper(i) dye-sensitised TiO2-based system for efficient dioxide by water vapor on mesoporous titania modified by bimetallic Au/Cu
light harvesting and photoconversion of CO2 into hydrocarbon fuel, Dalton nanostructures, Theory Exp. Chem. 50 (1) (2014) 53–58.
Trans. 41 (32) (2012) 9594–9597. [163] T. Ohno, et al., Preparation of S-doped TiO2 photocatalysts and their
[130] M. Anpo, Preparation, characterization, and reactivities of highly functional photocatalytic activities under visible light, Appl. Catal. A 265 (1) (2004) 115–
titanium oxide-based photocatalysts able to operate under UV–visible light 121.
irradiation: approaches in realizing high efficiency in the use of visible light, [164] K. Kondo, et al., Development of highly efficient sulfur-doped TiO2
Bull. Chem. Soc. Jpn. 77 (8) (2004) 1427–1442. photocatalysts hybridized with graphitic carbon nitride, Appl. Catal. B 142
[131] C. Comninellis, G. Chen, Electrochemistry for the Environment, Springer, (2013) 362–367.
2010, 2015. [165] J. Kim, W. Choi, H. Park, Effects of TiO2 surface fluorination on photocatalytic
[132] H. Shon, et al., Visible light responsive titanium dioxide (TiO2), J. Korean Ind. degradation of methylene blue and humic acid, Res. Chem. Intermed. 36 (2)
Eng. Chem. 19 (1) (2008) 1–16. (2010) 127–140.
[133] Y. Ma, et al., Titanium dioxide-based nanomaterials for photocatalytic fuel [166] C. Di Valentin, et al., N-doped TiO2: theory and experiment, Chem. Phys. 339
generations, Chem. Rev. 114 (19) (2014) 9987–10043. (1) (2007) 44–56.
[134] W. Sigmund, H. El-Shall, O. Dinesh Shah, M. Brij Moudgil, Particulate Systems [167] C. Di Valentin, et al., Characterization of paramagnetic species in N-doped
in Nano- and Biotechnologies, CRC Press, Boca Raton, 2008. TiO2 powders by EPR spectroscopy and DFT calculations, J. Phys. Chem. B 109
[135] H. Fujii, et al., Preparation and photocatalytic activities of a semiconductor (23) (2005) 11414–11419.
composite of CdS embedded in a TiO2 gel as a stable oxide semiconducting [168] N. Serpone, et al., Spectroscopic, photoconductivity, and photocatalytic
matrix, J. Mol. Catal. A: Chem. 129 (1) (1998) 61–68. studies of TiO2 colloids: naked and with the lattice doped with Cr3+, Fe3+, and
[136] H.C. Liang, X.Z. Li, J. Nowotny, Photocatalytical properties of TiO2 nanotubes, V5+ cations, Langmuir 10 (3) (1994) 643–652.
Solid State Phenom. 162 (2010) 295–328. [169] S. Rehman, et al., Strategies of making TiO2 and ZnO visible light active, J.
[137] G. Liu, et al., Titania-based photocatalysts–crystal growth, doping and Hazard. Mater. 170 (2–3) (2009) 560–569.
heterostructuring, J. Mater. Chem. 20 (5) (2010) 831–843. [170] T. Morikawa, Y. Irokawa, T. Ohwaki, Enhanced photocatalytic activity of
[138] D.A. Neamen, Semiconductor Physics and Devices: Basic Principles, McGraw TiO2xNx loaded with copper ions under visible light irradiation, Appl. Catal.
Hill, Singapore, 2012. A 314 (1) (2006) 123–127.
[139] M.F. Ashby, R.W. Messler, R. Asthana, E.P. Furlani, R.E. Smallman, A.H.W. Ngan, [171] T. Morikawa, et al., Visible-light-induced photocatalytic oxidation of
R.J. Crawford, N. Mills, Engineering Materials and Processes Desk Reference, carboxylic acids and aldehydes over N-doped TiO2 loaded with Fe, Cu or Pt,
Oxford, Butterworth-Heinemann, 2009. Appl. Catal. B 83 (1) (2008) 56–62.
[140] B.D. Fahlman, Materials Chemistry, Springer, Netherlands, 2008. [172] L. Huang, C. Sun, Y. Liu, Pt/N-codoped TiO2 nanotubes and its photocatalytic
[141] B. Pajot, B. Clerjaud, Optical Absorption of Impurities and Defects in activity under visible light, Appl. Surf. Sci. 253 (17) (2007) 7029–7035.
Semiconducting Crystals, vol. 169, Springer, Netherlands, 2013. [173] Y. Cong, et al., Preparation, photocatalytic activity, and mechanism of nano-
[142] Seebauer, G. Edmund, Kratzer, C. Meredith, Charged Semiconductor Defects: TiO2Co-doped with nitrogen and iron(III), J. Phys. Chem. C 111 (28) (2007)
Structure, Thermodynamics and Diffusion, Springer, Netherlands, 2009. 10618–10623.
[143] D.A. Hanaor, C.C. Sorrell, Review of the anatase to rutile phase [174] Q. Li, et al., Enhanced visible-light photocatalytic degradation of humic acid
transformation, J. Mater. Sci. 46 (4) (2011) 855–874. by palladium-modified nitrogen-doped titanium oxide, J. Am. Ceram. Soc. 90
[144] M. Anpo, Utilization of TiO2 photocatalysts in green chemistry, Pure Appl. (12) (2007) 3863–3868.
Chem. 72 (7) (2000) 1265–1270. [175] L. Zhou, et al., Preparation and characterization of N–I co-doped nanocrystal
[145] M. Anpo, et al., The design and development of second-generation titanium anatase TiO2 with enhanced photocatalytic activity under visible-light
oxide photocatalysts able to operate under visible light irradiation by irradiation, Mater. Chem. Phys. 117 (2) (2009) 522–527.
applying a metal ion-implantation method, Res. Chem. Intermed. 27 (4) [176] X. Yang, et al., Highly visible-light active C-and V-doped TiO2 for degradation
(2001) 459–467. of acetaldehyde, J. Catal. 252 (2) (2007) 296–302.
[146] M. Anpo, S. Dohshi, M. Takeuchi, Preparation of Ti/B binary oxide thin films by [177] X. Yang, et al., Mixed phase titania nanocomposite codoped with metallic
the ionized cluster beam (ICB) method: their photocatalytic reactivity and silver and vanadium oxide: new efficient photocatalyst for dye degradation, J.
photoinduced superhydrophilic properties, J. Ceram. Process. Res. 3 (4) Hazard. Mater. 175 (1) (2010) 429–438.
(2002) 258–260. [178] P. Richardson, et al., RETRACTED: manganese-and copper-doped titania
[147] M. Anpo, M. Takeuchi, The design and development of highly reactive nanocomposites for the photocatalytic reduction of carbon dioxide into
titanium oxide photocatalysts operating under visible light irradiation, J. methanol, Appl. Catal. B 126 (2012) 200–207.
Catal. 216 (1) (2003) 505–516. [179] P. Richardson, et al., RETRACTED: heterogeneous photo-enhanced conversion
[148] T. Umebayashi, et al., Analysis of electronic structures of 3d transition metal- of carbon dioxide to formic acid with copper-and gallium-doped titania
doped TiO2 based on band calculations, J. Phys. Chem. Solids 63 (10) (2002) nanocomposites, Appl. Catal. B 132 (2013) 408–415.
1909–1920. [180] X. Li, et al., Photocatalytic reduction of CO2 over noble metal-loaded and
[149] X. Nie, S. Zhuo, G. Maeng, K. Sohlberg, Doping of TiO2 polymorphs for altered nitrogen-doped mesoporous TiO2, Appl. Catal. A 429 (2012) 31–38.
optical and photocatalytic properties, Int. J. Photoenergy (2009) 1–22. [181] Q. Zhang, et al., Copper and iodine co-modified TiO2 nanoparticles for
[150] K. Kocí, et al., Effect of silver doping on the TiO2 for photocatalytic reduction improved activity of CO2 photoreduction with water vapor, Appl. Catal. B 123
of CO2, Appl. Catal. B 96 (3–4) (2010) 239–244. (2012) 257–264.
[151] B. Xin, et al., Study on the mechanisms of photoinduced carriers separation [182] B.-J. Liu, T. Torimoto, H. Yoneyama, Photocatalytic reduction of CO2 using
and recombination for Fe3+–TiO2 photocatalysts, Appl. Surf. Sci. 253 (9) surface-modified CdS photocatalysts in organic solvents, J. Photochem.
(2007) 4390–4395. Photobiol. A 113 (1) (1998) 93–97.
[152] A. Nishimura, G. Mitsui, M. Hirota, E. Hu, CO2 reforming performance and [183] Z.H. Zhao, et al., Optimal design and preparation of titania-supported CoPc
visible light responsibility of Cr-doped TiO2 prepared by sol–gel and dip- using sol–gel for the photo-reduction of CO2, Chem. Eng. J. 151 (1–3) (2009)
coating method, Int. J. Chem. Eng. (2010) 1–9. 134–140.
[153] J.C.S. Wu, H.M. Lin, Photo reduction of CO2 to methanol via TiO2 [184] H.-Y. Wu, H. Bai, J.C. Wu, Photocatalytic Reduction of CO2 using Ti-MCM-
photocatalyst, Int. J. Photoenergy 7 (3) (2005) 115–119. 41 photocatalysts in monoethanolamine solution for methane production,
[154] J. Wu, et al., Application of optical-fiber photoreactor for CO2 photocatalytic Ind. Eng. Chem. Res. 53 (28) (2014) 11221–11227.
reduction, Top. Catal. 47 (3) (2008) 131–136. [185] A. Dhakshinamoorthy, et al., Photocatalytic CO2 reduction by TiO2 and related
[155] O. Ola, et al., Performance comparison of CO2 conversion in slurry and titanium containing solids, Energy Environ. Sci. 5 (11) (2012) 9217–9233.
monolith photoreactors using Pd and Rh–TiO2 catalyst under ultraviolet [186] S. Kaneco, et al., Photocatalytic reduction of high pressure carbon dioxide
irradiation, Appl. Catal. B 126 (2012) 172–179. using TiO2 powders with a positive hole scavenger, J. Photochem. Photobiol. A
[156] R. Asahi, T. Morikawa, Nitrogen complex species and its chemical nature in 115 (3) (1998) 223–226.
TiO2 for visible-light sensitized photocatalysis, Chem. Phys. 339 (1–3) (2007) [187] Z. Zhao, et al., Optimal design and preparation of titania-supported CoPc
57–63. using sol–gel for the photo-reduction of CO2, Chem. Eng. J. 151 (1) (2009)
[157] M.M. Ren, Inverse opal titania on optical fiber for the photoreduction of CO2 134–140.
to CH3OH, Int. J. Chem. Reactor Eng. 7 (2009) . [188] K. Ko9cí, et al., Effect of temperature, pressure and volume of reacting phase on
[158] M. Schiavello, Heterogeneous Photocatalysis, John Wiley & Sons, Chichester, photocatalytic CO2 reduction on suspended nanocrystalline TiO2, Collect.
1997. Czech. Chem. Commun. 73 (8) (2008) 1192–1204.
O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42 41

[189] J.-M. Herrmann, Heterogeneous photocatalysis: fundamentals and [223] H. Lin, K. Valsaraj, Development of an optical fiber monolith reactor for
applications to the removal of various types of aqueous pollutants, Catal. photocatalytic wastewater Treatment, J. Appl. Electrochem. 35 (7) (2005)
Today 53 (1) (1999) 115–129. 699–708.
[190] H. Yamashita, et al., Photocatalytic synthesis of CH4 and CH3 OH from CO2 and [224] K. Nakata, A. Fujishima, TiO2 photocatalysis: design and applications, J.
H2O on highly dispersed active titanium oxide catalysts, Energy Convers. Photochem. Photobiol. C 13 (3) (2012) 169–189.
Manage. 36 (6) (1995) 617–620. [225] M. Tahir, N.S. Amin, Photocatalytic CO2 reduction with H2O vapors using
[191] F. Saladin, I. Alxneit, Temperature dependence of the photochemical montmorillonite/TiO2 supported microchannel monolith photoreactor,
reduction of CO2 in the presence of H2O at the solid/gas interface of TiO2, J. Chem. Eng. J. 230 (2013) 314–327.
Chem. Soc. Faraday Trans. 93 (23) (1997) 4159–4163. [226] M. Tahir, N.S. Amin, Photocatalytic CO2 reduction and kinetic study over In/
[192] G. Guan, et al., Photoreduction of carbon dioxide with water over K2Ti6O13 TiO2 nanoparticles supported microchannel monolith photoreactor, Appl.
photocatalyst combined with Cu/ZnO catalyst under concentrated sunlight, Catal. A 467 (2013) 483–496.
Appl. Catal. A 249 (1) (2003) 11–18. [227] M. Singh, I. Salvadó-Estivill, G. Li Puma, Radiation field optimization in
[193] Z.Y. Wang, et al., CO2 photoreduction using NiO/InTaO4 in optical-fiber photocatalytic monolith reactors for air treatment, AIChE J. 53 (3) (2007)
reactor for renewable energy, Appl. Catal. A 380 (1–2) (2010) 172–177. 678–686.
[194] S. Kaneco, et al., Photocatalytic reduction of CO2 using TiO2 powders in [228] M. Hossain, et al., Three-dimensional developing flow model for
supercritical fluid CO2, Energy 24 (1) (1999) 21–30. photocatalytic monolith reactors, AIChE J. 45 (6) (1999) 1309–1321.
[195] T. Mizuno, et al., Effect of CO2 pressure on photocatalytic reduction of CO2 [229] P.-Y. Liou, et al., Photocatalytic CO2 reduction using an internally illuminated
using TiO2 in aqueous solutions, J. Photochem. Photobiol. A 98 (1) (1996) 87– monolith photoreactor, Energy Environ Sci 4 (4) (2011) 1487–1494.
90. [230] O. Ola, M.M. Maroto-Valer, Copper based TiO2 honeycomb monoliths for CO2
[196] D. Erickson, D. Sinton, D. Psaltis, Optofluidics for energy applications, Nat. photoreduction, Catal. Sci. Tech. 4 (6) (2014) 1631–1637.
Photonics 5 (10) (2011) 583–590. [231] M.T. Merajin, et al., Photocatalytic conversion of greenhouse gases (CO2 and
[197] R. Howe, Recent developments in photocatalysis, Dev. Chem. Eng. Mineral CH4) to high value products using TiO2 nanoparticles supported on stainless
Process. 6 (1–2) (1998) 55–84. steel webnet, J. Taiwan Inst. Chem. Eng. 44 (2) (2013) 239–246.
[198] K. Koci, et al., Effect of TiO2 particle size on the photocatalytic reduction of [232] A. Nishimura, et al., CO2 reforming into fuel using TiO2 photocatalyst and gas
CO2, Appl. Catal. B 89 (3–4) (2009) 494–502. separation membrane, Catal. Today 148 (3–4) (2009) 341–349.
[199] C.B. Almquist, P. Biswas, Role of synthesis method and particle size of [233] P. Pathak, et al., Metal-coated nanoscale TiO2 catalysts for enhanced CO2
nanostructured TiO2 on its photoactivity, J. Catal. 212 (2) (2002) 145–156. photoreduction, Green Chem. 7 (9) (2005) 667–670.
[200] Z. Zhang, et al., Role of particle size in nanocrystalline TiO2-based [234] A. Cybula, M. Klein, A. Zielin  ska-Jurek, M. Janczarek, A. Zaleska, Carbon
photocatalysts, J. Phys. Chem. B 102 (52) (1998) 10871–10878. dioxide photoconversion. The effect of titanium dioxide immobilization
[201] A.N. Banerjee, The design, fabrication, and photocatalytic utility of conditions and photocatalyst type, Physicochem. Prob. Mineral Process. 48
nanostructured semiconductors: focus on TiO2-based nanostructures, (1) (2012) 159–167.
Nanotechnol. Sci. Appl. 4 (2011) 35. [235] M.N. Chong, et al., Recent developments in photocatalytic water treatment
[202] R.J. Braham, A.T. Harris, Review of major design and scale-up considerations technology: a review, Water Res. 44 (10) (2010) 2997–3027.
for solar photocatalytic reactors, Ind. Eng. Chem. Res. 48 (19) (2009) 8890– [236] M. Xing, F. Shen, B. Qiu, J. Zhang, Highly-dispersed boron-doped graphene
8905. nanosheets loaded with TiO2 nanoparticles for enhancing CO2
[203] Y. Shioya, et al., Synthesis of transparent Ti-containing mesoporous silica thin photoreduction, Sci. Rep. 4 (6341) (2014) 1–6.
film materials and their unique photocatalytic activity for the reduction of [237] Q. Zhang, et al., Photoreduction of carbon dioxide by graphene–titania and
CO2 with H2O, Appl. Catal. A 254 (2) (2003) 251–259. zeolite–titania composites under low-intensity irradiation, Mater. Sci.
[204] K. Ikeue, et al., Photocatalytic reduction of CO2 with H2O on Ti-beta zeolite Semicond. Process. 30 (2015) 162–168.
photocatalysts: Effect of the hydrophobic and hydrophilic properties, J. Phys. [238] E.S. Baeissa, Green synthesis of methanol by photocatalytic reduction of CO2
Chem. B 105 (35) (2001) 8350–8355. under visible light using a graphene and tourmaline co-doped titania
[205] O. Kozák, et al., Preparation and characterization of ZnS nanoparticles nanocomposites, Ceram. Int. 40 (8) (2014) 12431–12438.
deposited on montmorillonite, J. Colloid Interface Sci. 352 (2) (2010) 244– [239] L.-L. Tan, et al., Noble metal modified reduced graphene oxide/TiO2 ternary
251. nanostructures for efficient visible-light-driven photoreduction of carbon
[206] A.Y. Shan, T.I.M. Ghazi, S.A. Rashid, Immobilisation of titanium dioxide onto dioxide into methane, Appl. Catal. B 166–167 (2015) 251–259.
supporting materials in heterogeneous photocatalysis: a review, Appl. Catal. [240] M.M. Gui, S.-P. Chai, A.R. Mohamed, Modification of MWCNT@TiO2 core–shell
A 389 (1–2) (2010) 1–8. nanocomposites with transition metal oxide dopants for photoreduction of
[207] M. Dijkstra, et al., Comparison of the efficiency of immobilized and carbon dioxide into methane, Appl. Surf. Sci. 319 (2014) 37–43.
suspended systems in photocatalytic degradation, Catal. Today 66 (2) (2001) [241] J. Low, et al., Two-dimensional layered composite photocatalysts, Chem.
487–494. Commun. 50 (74) (2014) 10768–10777.
[208] Y.-H. Yu, et al., Photocatalytic NO reduction with C3H8 using a monolith [242] W. Fan, Q. Zhang, Y. Wang, Semiconductor-based nanocomposites for
photoreactor, Catal. Today 174 (1) (2011) 141–147. photocatalytic H2 production and CO2 conversion, Phys. Chem. Chem. Phys.
[209] O. Ozcan, et al., Dye sensitized CO2 reduction over pure and platinized TiO2, 15 (8) (2013) 2632–2649.
Top. Catal. 44 (4) (2007) 523–528. [243] W. Tu, et al., Robust hollow spheres consisting of alternating titania
[210] E.G. Look, H.D. Gafney, Photocatalyzed conversion of CO2 to CH4: an excited- nanosheets and graphene nanosheets with high photocatalytic activity for
state acid–base mechanism, J. Phys. Chem. A 117 (47) (2013) 12268–12279. CO2 conversion into renewable fuels, Adv. Funct. Mater. 22 (6) (2012) 1215–
[211] L. Liu, et al., Porous microspheres of MgO-patched TiO2 for CO2 1221.
photoreduction with H2O vapor: temperature-dependent activity and [244] M.M. Gui, et al., Enhanced visible light responsive MWCNT/TiO2 core–shell
stability, Chem. Commun. 49 (35) (2013) 3664–3666. nanocomposites as the potential photocatalyst for reduction of CO2 into
[212] J.Z. Tan, et al., Photoreduction of CO2 using copper-decorated TiO2 nanorod methane, Sol. Energy Mater. Sol. Cells 122 (2014) 183–189.
films with localized surface plasmon behavior, Chem. Phys. Lett. 531 (2012) [245] K.L. Choy, Chemical vapour deposition of coatings, Prog. Mater. Sci. 48 (2)
149–154. (2003) 57–170.
[213] W.-N. Wang, et al., Size and structure matter: enhanced CO2 photoreduction [246] V. Galindo, et al., High quality YBa2Cu3O7d/PrBa2Cu3O7d multilayers
efficiency by size-resolved ultrafine Pt nanoparticles on TiO2 single crystals, J. grown by pulsed injection MOCVD, J. Cryst. Growth 208 (1) (2000)
Am. Chem. Soc. 134 (27) (2012) 11276–11281. 357–364.
[214] D. Chen, F. Li, A.K. Ray, Effect of mass transfer and catalyst layer thickness on [247] C. Ying, D. Hao, W. Lishi, Doped-TiO2 photocatalysts and synthesis methods to
photocatalytic reaction, AIChE J. 46 (5) (2000) 1034–1045. prepare TiO2 films, J. Mater. Sci. Technol. 24 (5) (2008) 675–689.
[215] W. Choi, et al., Investigation on TiO2-coated optical fibers for gas-phase [248] X. Liu, et al., An improvement on sol–gel method for preparing ultrafine and
photocatalytic oxidation of acetone, Appl. Catal. B 31 (3) (2001) 209–220. crystallized titania powder, Mater. Sci. Eng. 289 (1–2) (2000) 241–245.
[216] A. Danion, et al., Characterization and study of a single-TiO2-coated optical [249] V. Meille, Review on methods to deposit catalysts on structured surfaces,
fiber reactor, Appl. Catal. B 52 (3) (2004) 213–223. Appl. Catal. A 315 (2006) 1–17.
[217] N.J. Peill, M.R. Hoffmann, Development and optimization of a TiO2-coated [250] U.G. Akpan, B.H. Hameed, The advancements in sol–gel method of doped-
fiber-optic cable reactor: photocatalytic degradation of 4-chlorophenol, TiO2 photocatalysts, Appl. Catal. A 375 (1) (2010) 1–11.
Environ. Sci. Technol. 29 (12) (1995) 2974–2981. [251] C.-C. Wang, J.Y. Ying, Sol–gel synthesis and hydrothermal processing of
[218] J. Xu, et al., Photocatalytic activity on TiO2-coated side-glowing optical fiber anatase and rutile titania nanocrystals, Chem. Mater. 11 (11) (1999) 3113–
reactor under solar light, J. Photochem. Photobiol. A 199 (2–3) (2008) 165– 3120.
169. [252] S. Sivakumar, et al., Nanoporous titania–alumina mixed oxides—an alkoxide
[219] T.V. Nguyen, J.C.S. Yu, Photoreduction of CO2 in an optical-fiber photoreactor: free sol–gel synthesis, Mater. Lett. 58 (21) (2004) 2664–2669.
effects of metals addition and catalyst carrier, Appl. Catal. A 335 (1) (2008) [253] Y. Li, T.J. White, S.H. Lim, Low-temperature synthesis and microstructural
112–120. control of titania nano-particles, J. Solid State Chem. 177 (4–5) (2004) 1372–
[220] P. Du, et al., A novel photocatalytic monolith reactor for multiphase 1381.
heterogeneous photocatalysis, Appl. Catal. A. 334 (1–2) (2008) 119–128. [254] J.C.S. Wu, I.H. Tseng, W.-C. Chang, Synthesis of titania-supported copper
[221] A.E. László Guczi, Catalysis for Alternative Energy Generation, Springer, New nanoparticles via refined alkoxide Sol–gel process, J. Nanoparticle Res. 3 (2)
York, 2012. (2001) 113–118.
[222] R.E. Marinangeli, D.F. Ollis, Photoassisted heterogeneous catalysis with [255] C. Su, B.-Y. Hong, C.-M. Tseng, Sol–gel preparation and photocatalysis of
optical fibers: I. Isolated single fiber, AIChE J. 23 (4) (1977) 415–426. titanium dioxide, Catal. Today 96 (3) (2004) 119–126.
42 O. Ola, M.M. Maroto-Valer / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 24 (2015) 16–42

[256] A. Nishimura, G. Mitsui, K. Nakamura, M. Hirota, E. Hu, CO2 reforming [271] T. Ohno, et al., Photocatalytic reduction of CO2 over a hybrid photocatalyst
characteristics under visible light response of Cr-or Ag-doped TiO2 prepared composed of WO3 and graphitic carbon nitride (g-C3N4) under visible light, J.
by sol–gel and dip-coating process, Int. J. Photoenergy (2012) 1–12. CO2 Util. 6 (2014) 17–25.
[257] M. Langlet, et al., Liquid phase processing and thin film deposition of titania [272] C.W. Lee, R.A. Kourounioti, J.C. Wu, E. Murchie, M. Maroto-Valer, O.E. Jensen,
nanocrystallites for photocatalytic applications on thermally sensitive A. Ruban, Photocatalytic conversion of CO2 to hydrocarbons by light-
substrates, J. Mater. Sci. 38 (19) (2003) 3945–3953. harvesting complex assisted Rh-doped TiO2 photocatalyst, J. CO2 Util. 5
[258] J. Zhang, et al., Importance of the relationship between surface phases and (2014) 33–40.
photocatalytic activity of TiO2, Angew. Chem. Int. Ed. 47 (9) (2008) 1766– [273] P. Silija, Z. Yaakob, V. Suraja, N.N. Binitha, Z.S. Akmal, An enthusiastic glance in
1769. to the visible responsive photocatalysts for energy production and pollutant
[259] B. Vijayan, et al., Effect of calcination temperature on the photocatalytic removal, with special emphasis on titania, Int. J. Photoenergy (2012) 1–19.
reduction and oxidation processes of hydrothermally synthesized titania [274] K. Nishida, et al., In-situ monitoring of PE-CVD growth of TiO2 films with laser
nanotubes, J. Phys. Chem. C 114 (30) (2010) 12994–13002. Raman spectroscopy, Appl. Surf. Sci. 159 (2000) 143–148.
[260] K.L. Schulte, P.A. DeSario, K.A. Gray, Effect of crystal phase composition on the [275] C. Perego, S. Peratello, Experimental methods in catalytic kinetics, Catal.
reductive and oxidative abilities of TiO2 nanotubes under UV and visible light, Today 52 (2–3) (1999) 133–145.
Appl. Catal. B 97 (3) (2010) 354–360. [276] T. Van Gerven, et al., A review of intensification of photocatalytic processes,
[261] J. Nair, et al., Microstructure and phase transformation behavior of doped Chem. Eng. Process. 46 (9) (2007) 781–789.
nanostructured titania, Mater. Res. Bull. 34 (8) (1999) 1275–1290. [277] H. Lin, An optical fiber monolith reactor for photocatalytic wastewater
[262] G. Liu, et al., Engineering TiO2 nanomaterials for CO2 conversion/solar fuels, treatment, AIChE J. 52 (6) (2006) 2271–2280.
Sol. Energy Mater. Sol. Cells 105 (2012) 53–68. [278] R.-D. Sun, et al., TiO2-coated optical fiber bundles used as a photocatalytic
[263] J.T. Carneiro, et al., How phase composition influences optoelectronic and filter for decomposition of gaseous organic compounds, J. Photochem.
photocatalytic properties of TiO2, J. Phys. Chem. C 115 (5) (2011) Photobiol. A 136 (1–2) (2000) 111–116.
2211–2217. [279] M. Bouchy, O. Zahraa, Photocatalytic reactors, Int. J. Photoenergy 5 (3) (2003)
[264] C.C. Yang, et al., Artificial photosynthesis over crystalline TiO2-based 191–197.
catalysts: fact or fiction? J. Am. Chem. Soc. 132 (24) (2010) 8398–8406. [280] P.S. Mukherjee, Ray, Major challenges in the design of a large-scale
[265] A. Cybula, M. Klein, A. Zaleska, Methane formation over TiO2-based photocatalytic reactor for water treatment, Chem. Eng. Technol. 22 (3) (1999)
photocatalysts: reaction pathways, Appl. Catal. B 164 (2015) 433–442. 253.
[266] T. Yui, et al., Photochemical reduction of CO2 using TiO2: effects of organic [281] A.K. Ray, Design, modelling and experimentation of a new large-scale
adsorbates on TiO2 and deposition of Pd onto TiO2, ACS Appl. Mater. photocatalytic reactor for water treatment, Chem. Eng. Sci. 54 (15) (1999)
Interfaces 3 (7) (2011) 2594–2600. 3113–3125.
[267] K. Teramura, et al., Photocatalytic conversion of CO2 in water over [282] J.-J. Wang, et al., Photocatalytic reduction of CO2 to energy products using Cu–
layered double hydroxides, Angew. Chem. Int. Ed. 51 (32) (2012) 8008– TiO2/ZSM-5 and Co–TiO2/ZSM-5 under low energy irradiation, Catal.
8011. Commun. 59 (2015) 69–72.
[268] B.-J. Liu, et al., Effect of solvents on photocatalytic reduction of carbon dioxide [283] B. Sánchez, et al., Influence of temperature on gas-phase photo-assisted
using TiO2 nanocrystal photocatalyst embedded in SiO2 matrices, J. mineralization of TCE using tubular and monolithic catalysts, Catal. Today 54
Photochem. Photobiol. A 108 (2–3) (1997) 187–192. (2–3) (1999) 369–377.
[269] Y. Fu, et al., An amine-functionalized titanium metal–organic framework [284] J.T. Carneiro, et al., An internally illuminated monolith reactor: pros and cons
photocatalyst with visible-light-induced activity for CO2 reduction, Angew. relative to a slurry reactor, Catal. Today 147 (Suppl. 1) (2009) S324–S329.
Chem. Int. Ed. 51 (14) (2012) 3364–3367. [285] W. Wang, The light transmission and distribution in an optical fiber coated
[270] T. Ohno, et al., Photocatalytic reduction of CO2 over exposed-crystal-face- with TiO2 particles, Chemosphere 50 (8) (2003) 999–1006.
controlled TiO2 nanorod having a brookite phase with co-catalyst loading, [286] J.C.S. Wu, H.-M. Lin, C.-L. Lai, Photo reduction of CO2 to methanol using
Appl. Catal. B 152 (2014) 309–316. optical-fiber photoreactor, Appl. Catal. A 296 (2) (2005) 194–200.

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy