Accepted Manuscript A review on photo-thermal catalytic conversion of carbon dioxide Ee Teng Kho, Tze Hao Tan, Emma Lovell, Roong Jien Wong, Jason Scott, Rose Amal PII:
S2468-0257(17)30064-X
DOI:
10.1016/j.gee.2017.06.003
Reference:
GEE 73
To appear in:
Green Energy and Environment
Received Date: 3 April 2017 Revised Date:
5 June 2017
Accepted Date: 5 June 2017
Please cite this article as: E.T. Kho, T.H. Tan, E. Lovell, R.J. Wong, J. Scott, R. Amal, A review on photo-thermal catalytic conversion of carbon dioxide, Green Energy & Environment (2017), doi: 10.1016/ j.gee.2017.06.003. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
RI PT
ACCEPTED MANUSCRIPT
Energy-harvesting from the sun via concentrated solar irradiation or localised plasmonic
AC C
EP
TE D
M AN U
SC
heating for catalytic conversion of carbon dioxide
ACCEPTED MANUSCRIPT Article Type: Mini Review
Title:
RI PT
A review on photo-thermal catalytic conversion of carbon dioxide Ee Teng Kho, Tze Hao Tan, Emma Lovell, Roong Jien Wong, Jason Scott*, and Rose Amal*
SC
Particles and Catalysis Research Group School of Chemical Engineering
Sydney 2052 New South Wales, Australia. Email:
[email protected]
AC C
EP
TE D
[email protected]
M AN U
The University of New South Wales
ACCEPTED MANUSCRIPT
Abstract The conversion of carbon dioxide into value-added products is of great industrial and environmental interest. However, as carbon dioxide is relatively stable, the input energy required for this conversion is a significant limiting factor in the system’s performance. By
RI PT
utilising energy from the sun, through a range of key routes, this limitation can be overcome. In this review, we present a comprehensive and critical overview of the potential routes to harvest the sun’s energy, primarily through solar-thermal technologies and plasmonic
SC
resonance effects. Focusing on the localised heating approach, this review shortlists and compares viable catalysts for the photo-thermal catalytic conversion of carbon dioxide.
M AN U
Further, the pathways and potential products of different carbon dioxide conversion routes are outlined with the reverse water gas shift, methanation, and methanol synthesis being of key interest. Finally, the challenges in implementing such systems and the outlook to the
Keywords
TE D
future are detailed.
AC C
EP
Carbon dioxide conversion; photo-thermal; plasmonic catalysis; solar thermal
1
ACCEPTED MANUSCRIPT
Table of Contents Abstract ...................................................................................................................................... 1 Keywords ................................................................................................................................... 1 Table of Contents ....................................................................................................................... 2 Introduction ........................................................................................................................ 3
2.
Overview on photo-thermal catalysis ................................................................................ 4
RI PT
1.
Motivation for photo-thermal catalysis ....................................................................... 4
2.2.
Heat-harvesting for photo-thermal catalysis ............................................................... 6
SC
2.1.
Solar thermal technologies ................................................................................... 6
2.2.2.
Plasmonic resonance effect .................................................................................. 9
2.2.3.
Evaluating the relative strengths and weaknesses of solar thermal vs. plasmonic
M AN U
2.2.1.
technologies for photo-thermal CO2 conversion.............................................................. 13 Overview of common CO2 conversion products and pathways ...................................... 14 3.1.
Reverse water-gas shift reaction ............................................................................... 14
3.2.
Methanation ............................................................................................................... 16
3.3.
Methanol synthesis .................................................................................................... 18
TE D
3.
Available candidates and opportunities in catalyst materials selection ........................... 19
5.
Challenges in the execution of catalytic photo-thermal CO2 reduction ........................... 25
6.
Conclusion and outlook ................................................................................................... 28
EP
4.
AC C
References ................................................................................................................................ 30
2
ACCEPTED MANUSCRIPT
1. Introduction It is undeniable that the global energy demand has an ever-growing thirst, driven primarily by our incessant craving for a better quality of life and the consequent urbanisation and economic growth. Despite rising environmental awareness and establishment of policies that
RI PT
have facilitated the diversification of energy sources to include more renewable alternatives, fossil fuels will continue to dominate as a primary source of energy [1]. One of the key consequences in having such a dependence on carbon-based energy sources is increasing
SC
carbon dioxide (CO2) emissions. Increasing emissions of this greenhouse gas and its subsequent accumulation in the atmosphere, due to its relative inertness, has led to great
caused to local ecosystems.
M AN U
concerns regarding the ensuing irregularities in global weather patterns and devastation
An appealing solution to the persistent fuel demand and the inevitable CO2 emissions is to
TE D
complete the carbon cycle via the recycling of CO2 into synthetic fuels. This is possible through the reduction of CO2. Nevertheless, one of the main hindrances to such a process is its intense energy consumption. Despite the overall exothermicity of reduction reactions, such
EP
as methanation and methanol synthesis, the relative stability of a CO2 molecule necessitates energy provision to initiate its conversion. The conventional industrial practice of deriving
AC C
process heat from fossil fuel incineration to fulfil the energy demand is not ideal as it further contributes to CO2 release. Therefore, in addition to studying effective pathways for CO2-tofuel conversions, it is also crucial that a renewable energy feed source, which is effective and suitable for such purpose, is accessible. As the Earth’s most abundant energy resource, the utilisation of sunlight as an alternative energy supply has become a popular field of scientific study. Several opportunities exist in its application to provide for the demand of CO2 reduction processes. While clean, solar energy-
3
ACCEPTED MANUSCRIPT powered chemical processes are more commonly associated with photocatalysis, the achievable conversion efficiencies at current stages of development are often unfeasible for large-scale operations as compared to the more industrially established thermal catalysis.
thermal catalytic CO2 reduction thus seem favourable.
RI PT
With solar irradiation comprising primarily both heat and light energy, prospects for photo-
In the ensuing sections, the possibility of harnessing heat and light from solar energy through solar thermal technologies and plasmonic resonance effects is outlined. Additionally, the
SC
application of such technologies for effective catalytic recycling of CO2, with particular
M AN U
reference to the strength and limitations of the differing routes, is reviewed. In addition to discussing the potentials of different candidate materials for simultaneously achieving successful solar absorption and CO2 reduction, the foreseen challenges ahead in the realisation of such a process are also assessed.
2.1.
TE D
2. Overview on photo-thermal catalysis Motivation for photo-thermal catalysis
EP
Thermal catalysis is one of the most common forms of catalysis, with conventional industrial processes such as the ammonia synthesis and reforming utilising heat input to compensate for
AC C
the endothermicity of these reactions. One of the key reasons for industrial dependence on thermally driven catalysis is the high efficiency and applicability for large scale processes. Nevertheless, the large energy demand is often provided for via the incineration of fossil fuels [2]. With growing concerns regarding the depletion of non-renewable fuel resources and rising awareness regarding environmental footprints, harnessing solar energy to drive energetically demanding catalytic processes is desirable. Aside from the photovoltaic transformation of sunlight into electricity to facilitate conversion, the direct activation of
4
ACCEPTED MANUSCRIPT catalytic materials through sunlight absorption is also possible. The work by Honda and Fujishima [3] on the use of a titanium dioxide semiconductor for water-splitting instigated a rapid expansion in the field of photocatalytic research. However, despite the energetically ideal concept of the direct translation of light into chemical energy, the practical application
RI PT
of photocatalytic technologies is greatly hindered by the lack of effective conversions, especially for industrially-scaled operations [4], with the difference in efficiencies often several orders of magnitude below the capabilities of thermal catalysis.
SC
This subsequently inspired the concept of light-for-heat in catalysis as an alternative to
M AN U
conventionally fossil fuel-powered systems. What distinguishes such photo-thermal catalysis from traditional thermal- or photo-catalysis is the utilisation of both the light and heat components of solar radiation, whereby heat is either obtained directly from or initiated by the light absorption. This thus enables the utilisation of a sustainable energy source (as with
thermal catalysis.
TE D
photocatalysis) all the while attaining effective catalytic conversions typically obtained for
The possible pathways to realising the concept are visualised in Figure 1. The ‘direct’
EP
procurement of heat from solar radiation is possible via the traditional concept of solar thermal heating. Solar thermal heating operates based on the principle of utilising a light-
AC C
absorbing material or infrastructure to capture heat from the visible or infrared spectrum of solar radiation. The trapped heat will then result in an elevation in temperature, the extent to which depends on the properties of the light-absorbing material and infrastructure design. Another pathway from which heat energy can be obtained from sunlight is through the plasmon resonance phenomenon. Although both these ‘direct’ and ‘indirect’ routes essentially rely on vibration at a subatomic scale, the surface plasmon resonance distinguishes itself from the conventional perception of heating. In addition to the
5
ACCEPTED MANUSCRIPT conventional heat energy production upon light absorption, this phenomenon allows the formation of intense, localised hotspots due to the resonance of surface plasmons of
M AN U
SC
RI PT
plasmonic materials upon exposure to radiation of certain frequencies.
2.2.
TE D
Figure 1. Sourcing heat from the sun for a photo-thermal catalytic system. Heating of the catalyst bed can be achieved either through directing concentrated solar irradiation onto the catalyst material or relying primarily on the localised plasmonic heating effects through surface plasmon excitation.
Heat-harvesting for photo-thermal catalysis
EP
There exists a range of routes to harvest heat from the sun for photo-thermal catalytic
AC C
conversions. These routes are examined in this Section, specifically in the context of solar thermal technologies and the plasmonic resonance effect. Finally, a critical evaluation of the application of such heat-harvesting approaches for carbon dioxide is presented.
2.2.1. Solar thermal technologies The sensation of heat felt upon exposure to sunlight crudely explains the concept of solar thermal heating. From residential rooftop water heating systems to solar furnaces capable of temperatures beyond 2300 K, records of solar thermal technology applications can be traced back to prehistoric times [5]. In heat exchanging applications for domestic purposes, whereby 6
ACCEPTED MANUSCRIPT temperature requirements are not as stringent (e.g. water-heating), the basic infrastructure comprises a collector upon which solar radiation is ‘gathered’ and reflected to a working fluid (either gaseous, liquid or molten salts) to conduct and/or circulate the absorbed energy. For higher temperature systems, the collector usually is designed to allow for light-concentrating
RI PT
capabilities for heat intensity amplification (Figure 2). Depending on the solar concentrating systems utilised, concentration ratios between the range of 30 – 100 (for a parabolic trough system to achieve 500 – 700 K temperatures, Figure 2a) and up to 5000 – 10000 (in a double-
SC
concentration system consisting of a heliostat field, a reflective tower and a ground receiver
EP
TE D
M AN U
capable of temperatures in excess of 1500 K, Figure 2d) are attainable [6].
Figure 2. Main high temperature solar concentrator system categories: (a) parabolic trough, (b) central power tower, (c) parabolic dish, and (d) double concentration. Adapted from [6].
AC C
To achieve effective heat trapping, two key qualities are required by the absorbing material: high solar absorption and low thermal emissivity. Materials with high solar absorption enable more effective heat generation whereas low thermal emissivity is essential to minimise subsequent heat loss through radiation. Therefore, in addition to the intuitive choice of using a dark-coloured heat absorbing material, much effort has been invested in the understanding of optical and thermal properties of materials for constructing solar absorber coatings such as paints, enamel, carbides, and cermets [7, 8].
7
ACCEPTED MANUSCRIPT One of the main appeals of the technique of heat harvesting is the extremely high temperatures attainable through the concentration of solar energy. As mentioned previously, temperatures beyond 1000 K are possible depending on the solar concentrator configuration. Such capability is subsequently taken advantage of for thermochemical conversions. These
RI PT
processes involve a redox cycle whereby firstly, high temperature (> 1500 K) treatment of a metal oxide (e.g. CeO2 [9-11] and ZnO [12-14]) is performed to obtain its reduced form. The ensuing step is then conducted at a slightly lower temperature whereby the reduced metal
SC
oxide is exposed to reactants such as CO2 (or H2O) to promote their reduction. Solar-aided reforming is another common conversion process that has been tested in a solar reactor
M AN U
configuration [4, 15-17]. A catalyst is usually utilised in order to lower the temperature requirements for conversion (≤ 1000 K). In this case, the catalyst can either be located in a reactor decoupled from the receiver or be included as part of the receiver where it is also
AC C
EP
TE D
exposed to irradiation (Figure 3).
Figure 3. Different configurations for a solar reactor: (left) reactor decoupled from solar receiver system, (middle) integrated tubular reactor/receiver, and (right) catalytically active solar absorber. Adapted from [4].
8
ACCEPTED MANUSCRIPT 2.2.2. Plasmonic resonance effect As opposed to focusing on infrastructure design, as is the case for solar thermal technologies, the plasmonic route places greater attention on developing the light-absorbing material itself. While both traditional photocatalysis and plasmonic catalysis rely on light-source activation,
RI PT
their similarities stop here. Conventional photocatalysis utilises light energy to overcome the bandgap energy difference of a semiconductor, subsequently generating active electrons and holes that are capable of a myriad of redox reactions depending on the reducing/oxidising
SC
potentials of the resulting charge carriers (i.e. the positions of the conduction/valence bands). Plasmonic catalysis, on the other hand, relies on the resonance of the oscillations of free
M AN U
electrons (i.e. plasmons) upon the absorption of incident electromagnetic radiation. Such resonance occurs when the delocalised electrons are supplemented with an energy wave, whose frequency corresponds to the characteristic resonant frequency of the metal
AC C
EP
TE D
nanoparticle (Figure 4).
Figure 4. Schematic of plasmon oscillation for a sphere, showing the displacement of the conduction electron charge cloud relative to the nuclei. Reprinted with permission from ref. [18]. Copyright (2003) American Chemical Society.
The radiation frequency required for such excitation depends on factors such as composition, shape, and size [19-24] as illustrated in Figure 5. The dependence of plasmonic response on particle composition and morphology can be explained by Mie theory, which has been readily exploited for theoretical analyses of particle optical properties [25]. Mie theory states that the 9
ACCEPTED MANUSCRIPT total extinction cross-section of a particle depends on its spherical volume, the dielectric constant of its surrounding environment, and the dielectric function of the particle. Factors such as composition, shape, and size influence the dielectric function of the particle. In polyhedral studies (Figure 5a), blue plasmonic shifts have been observed for particles with
RI PT
increasing number of facets, with spherical particles having the smallest plasmonic peak position [22]. In these studies, it was also noted that polyhedral particles may exhibit multiple surface plasmon resonances, leading to band broadening and the emergence of new discrete
SC
peaks [20]. Similarly, changing the composition of the particles through alloying can promote the formation of plasmonic peaks, notably as a mixture of the individual plasmonic
M AN U
components (Figure 5b) [23]. The dependency of plasmonic response on particle composition and morphology permits the flexible tuning of absorption wavelength based on the intended application.
In terms of material choices, gold, silver and copper are amongst the most commonly studied
TE D
metals for plasmonic applications due to their strong plasmonic responses [26]. Whilst nanosized gold and silver are popular candidates, nickel and copper have also garnered
EP
significant attention due to their relatively cheaper costs. These metals also have the advantage of having their plasmonic frequencies correspond to electromagnetic radiation
AC C
residing within the visible-near infrared region, which is in greater abundance within the solar spectrum as compared to the ultraviolet light, upon which most effective semiconductor photocatalysts are reliant on.
10
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
Figure 5. Influence of: (a) dimensions of silver nanoparticles over their peak absorption wavelength [23]; and (b) composition of silver-gold alloys over the spectral absorption profiles of the nanoparticles [24]. Reprinted with permission from ref. [24]. Copyright (2000) American Chemical Society.
For the purpose of photo-thermal catalysis, plasmonic resonance presents itself as an appealing alternative to conventional heating for reactions that require elevated temperatures. While the earlier studies into plasmonic-assisted reactions were mainly dedicated to their photochemical aspects, it was not until 2007 when Cao et al. [27] first looked into utilising thermoplasmonics for nanoscale material fabrication. Ni pigments have long been used for solar thermal heating despite not being previously deemed as plasmonic nanoparticles [28,
11
ACCEPTED MANUSCRIPT 29]. In contrast to conventional heating, where heat is provided externally, thermoplasmonics enables rapid and localised heating within the vicinity of the plasmonic nanoparticle [30]. In addition to the thermal energy supplement, reactant molecules in the vicinity of a
RI PT
plasmonic nanoparticle can experience an enhanced incident photon rate and/or photogenerated hot electron formation (Figure 6b-c [31]). Not to be confused with thermoplasmonics, photogenerated hot electrons have an athermal charge charrier distribution which cannot be described using a Fermi-Dirac distribution. Instead, their behaviours are
SC
more similar to electron-hole pairs in semiconductors [32]. The injection of hot electrons may
M AN U
occur directly from catalyst to reactant or through a semiconductor support interface, similar to that of a dye-sensitised photocatalyst [33]. By acting as both the optical absorption and catalytic sites, plasmonic photocatalysts have advantages over semiconductors due to their higher electron density and affinity to reactants. Consequently, in contrast to conventional photocatalysis, which typically provides yields in the micromole (per gram catalyst per hour)
AC C
EP
TE D
range, plasmonic catalysts are often capable of yields greater by orders of magnitude [34-36].
Figure 6. Possible mechanisms for plasmonic catalysis: (a) thermoplasmonic temperature increase for heat transfer to reactant within vicinity; (b) nearby reactant experiencing an apparent increase in incident photon rate (i.e. near field optical enhancement); (c) hot electron injection into nearby reactant; (d) plasmonic heat generation resulting in enhanced electronhole generation within a semiconductor for reactant molecule activation; (e) near field optical enhancement experienced by semiconductor, leading to enhanced electron-hole generation; and (f) hot electron injection from a plasmonic nanoparticle into the neighbouring semiconductor, which is then transferred into a reactant molecule nearby. Reproduced from ref. [31] with permission from The Royal Society of Chemistry.
12
ACCEPTED MANUSCRIPT 2.2.3. Evaluating the relative strengths and weaknesses of solar thermal vs. plasmonic technologies for photo-thermal CO2 conversion Aside from the capabilities and potential of the respective solar energy-harvesting techniques discussed above, it is important to acknowledge the relative limitations of the technologies
RI PT
for the design of a successful photo-thermal CO2 conversion system that is practical for real world applications. While laboratory-scale experiments often place a primary focus on aspects such as conversions and yields, industrial considerations such as scalability, cost
SC
effectiveness, and robustness of the system also have to be taken into account for successfully
M AN U
transferring novel technologies to commercial operations.
Solar thermal catalysis has the advantage of technical maturity with tests carried out in commercial scale-receiver/reactors since the 1980’s [2, 16, 37, 38]. While the tests have delivered promising results, the robustness of catalyst materials under the typically high
TE D
operating temperatures of solar thermal catalytic reactors is frequently cited to be the shortcoming of such systems with catalyst disintegration via sintering and/or the encapsulation of active phases being amongst the most commonly reported issues [16, 17, 39-
EP
41]. These occurrences largely stem from the fact that the high temperature reaction conditions place severe stress on the lifetime of a catalyst. Moreover, the need for a thermally
AC C
resistant reactor with high levels of technical precision, in addition to the large footprint requirements, leads to exorbitant costs associated with constructing solar thermal infrastructure.
On this basis, plasmonic catalysis appears to offer the advantage as the intense heating occurs locally, within the vicinity of the active nanoparticles. As a result, the demand for external heat supply to the system and overall thermal stress on the infrastructure is much less for plasmonic-based catalytic systems. In the case where heat is sourced from within the
13
ACCEPTED MANUSCRIPT plasmonic material itself, the robustness and stability of the material properties affecting plasmonic efficiencies, such as nanoparticle shape and size, over the course of prolonged usage (and therefore continuous exposure to heat), becomes even more crucial for maintaining consistent heating capabilities and conversion performance. As demonstrated by
RI PT
the available thermoplasmonic studies, strict requirements on the nanoparticles size (<5 nm), high incident light intensity, and/or extreme isolation of nanoparticles are currently the challenges faced for plasmonic catalysis [42]. At such a relative stage of infancy, most
SC
investigations remain invested in the development and understanding of materials and are yet
M AN U
to venture into feasibility studies for industrially scaled catalytic processes.
3. Overview of common CO2 conversion products and pathways Owing to the low CO2 conversion efficiency in CO2 reduction reactions via plasmonic catalysis, understanding CO2 chemistry is critical for improving the conversion efficiency.
TE D
CO2 is an apolar molecule due to the linear configuration of its two polar C=O bonds. Due to its relative inertness, the resulting highly positive Gibbs free energy means that its conversion
activate the CO2.
EP
is often not thermodynamically favourable. Consequently, a catalyst is often required to
AC C
Transforming CO2 into synthetic fuels can occur through different pathways into a myriad of C-based products. In this review, the focus will be placed upon its catalytic hydrogenation into CO, CH4, and CH3OH, which are often quintessential building blocks for the downstream production of fuels and chemical commodities.
3.1.
Reverse water-gas shift reaction
The reduction of carbon dioxide into carbon monoxide (Equation 1) is one of the more popular routes for CO2 conversion studies mainly due to its low hydrogen consumption (as
14
ACCEPTED MANUSCRIPT compared to methanation) and high equilibrium conversion rates [43]. Furthermore, the formation of CO forms an important first step for the subsequent synthesis of carbon-based chemicals and/or fuels [43-45]. ∆H = +41 kJ/mol
Equation 1
RI PT
CO2 + H2 ⇌ CO + H2O
The mechanisms suggested for reverse water-gas shift (RWGS) reaction fall into two main categories: (i) the redox pathway [46, 47] and (ii) the formate decomposition mechanism [48, 49]. The redox mechanism is most often described over a Cu-based catalyst, whereby the
SC
metallic Cu0 atoms are partially oxidised to form Cu2O and carbon monoxide (Equation 2)
M AN U
and the oxidised copper is later reduced by the hydrogen present in the system, reforming it to its metallic state accompanied by the formation of a water molecule (Equation 3). CO2(g) + 2Cu0(s) ⇌ CO(g) + Cu2O(s)
Equation 2
H2(g) + Cu2O(s) ⇌ H2O(g) + 2Cu0(s)
Equation 3
TE D
In contrast to the redox mechanism, whereby the hydrogen molecule does not actively interact with the carbon dioxide molecule, the formate decomposition pathway suggests that the hydrogen molecule first hydrogenates the carbon dioxide into a formate intermediate. It is
EP
then from this intermediate that the cleaving of the carbon-oxygen double bond occurs to release a CO molecule [50]. The formate-mediated route is also reported to occur for the
AC C
methanation pathway (discussed in Section 3.2), but has often been debated regarding the extent of its contribution toward product formation. An investigation by Goguet et al. [51] is one such example. Using a Pt/CeO2 catalyst for their in operando spectrokinetic investigation, they proposed that the Pt metal maintains the active (i.e. reduced) ceria surface via H2-spillover. The resulting vacancies on the ceria surface then bond with CO2 molecules to form surface carbonates, which are the main intermediates that facilitate the final formation of CO species. 15
ACCEPTED MANUSCRIPT
3.2.
Methanation
The production of methane from CO2 (Equation 4) is of great interest in certain geographical locations due to the ease in incorporating its distribution into existing infrastructure and
with the added bonus of regulating CO2 emissions. CO2 + 4·H2 ⇌ CH4 + 2·H2O
RI PT
networks [52]. Furthermore, it serves as an attractive solution for renewable energy storage
∆H = -165 kJ/mol
Equation 4
SC
In terms of the mechanistic pathway involved, claims from both experimental and theoretical studies again falls into two main domains: (i) direct hydrogenation of CO2 or (ii) CH4
M AN U
formation via a CO intermediate [53], with majority of the findings suggesting the latter process is the more probable methanation route.
First proposed by Bahr, CH4 formation via the CO pathway was supported by Peebles and Goodman in an investigation, which compared the methanation of CO and CO2 over a Ni(100)
TE D
surface [54]. They found the activation energy and reaction rate values for CH4 formation from CO2 and CH4 formation from CO were similar under identical reaction conditions. Later studies utilising spectroscopic methods, such as diffuse reflectance infrared spectroscopy
EP
(DRIFTS), and Fourier-transformed infrared spectroscopy (FTIR) coupled with mass spectroscopy, have observed both formate and CO intermediates on the catalyst surface [55-
AC C
57]. The ensuing interpretations, however, have led to different conclusions. Marwood et al. [55] proposed the formation of formate through a carbonate species, which was subsequently hydrogenated into an adsorbed CO species (COads) (illustrated in Figure 7), which has been suggested to be the rate determining step [56]. In contrast, an in situ DRIFTS study (using Rh/ γ-Al2O3) by Jacquemin et al. [57] reported instead that COads is a product of the direct dissociation of CO2 without mentioning the detection (and role) of formate species.
16
SC
RI PT
ACCEPTED MANUSCRIPT
Figure 7. Proposed CO2 methanation mechanism involving the formation of formate through a carbonate species. S = support, M = metal, I = metal-support interface [55].
M AN U
Experimental measurements coupled with computational methods have also been used to probe the role of intermediate species during CO2 hydrogenation on nickel. While it was acknowledged that both formate and CO are present under the reaction conditions commonly applied in the mechanistic studies of CO2 methanation, Vesselli et al. [58] demonstrated that
TE D
formate was purely a spectator species and that the reaction proceeded via the direct hydrogenation of the C-O bond of CO2. This conclusion arose mainly from the findings that formate has high reaction barriers for further hydrogenation for its transformation into CO.
EP
This agrees with the findings by Marwood et al. [55] who, despite proposing a different
AC C
methanation pathway, also reported the relative stability of the formed formate species, and further lends credibility to the argument of Goguet et al. [51] (discussed above) on the formation of CO not being dominated by the formate route. Ren et al. [59] performed density functional theory (DFT) calculations for the three mechanisms listed in Figure 8 on the Ni(111) surface with the energy involved in each step of the three paths being illustrated in Figure 9. A conclusion similar to Vesseli et al.’s was reached whereby a comparison of the energy barriers involved in the rate determining steps of each route revealed the mechanism
17
ACCEPTED MANUSCRIPT involving CO without direct participation of the formate species being the most energetically
SC
RI PT
plausible.
EP
TE D
M AN U
Figure 8. CO2 methanation mechanism: (a) via CO involving the formate species; (b) via CO without the participation of the formate species; and (c) direct hydrogenation of CO2 without the participation of a CO intermediate [59].
3.3.
AC C
Figure 9. Potential energy diagram for different CO2 methanation mechanisms. Path 1, 2, and 3 correspond to the pathways listed in Figure 8a, b, and c, respectively [59]. TS = transition state.
Methanol synthesis
One of the most apparent advantages methanol has over products such as CO and CH4 is its liquid state at normal atmospheric temperature and pressure. This is desirable as it provides a practical alternative for energy storage and transport. Furthermore, much like CO, it can also serve as an important starting constituent for olefins and aromatics [60-63]. One of the major challenges for the methanol synthesis process (Equation 5) is it is usually less selective than 18
ACCEPTED MANUSCRIPT methane and CO under low-pressure conditions. High pressures (50 – 100 bars) are often required to suppress CO formation via the RWGS pathway (Equation 1) [64]. ∆H = -49 kJ/mol
Equation 5
RI PT
CO2 +3·H2 ⇌ CH3OH + H2O
Similar to the Sabatier case, arguments on the CO2 to methanol reaction pathway are dominated by the formate route and the CO route factions. The pathway suggested for the latter case proposes that CO2 first undergoes RWGS-like hydrogenation before the CO
SC
formed is sequentially hydrogenated into formyl (HCO), formaldehyde (H2CO), methoxy
M AN U
(H3CO), and finally methanol [65]. The study by Grabow and Mavrikakis [63] suggested the formate pathway being more likely, whereby the hydrogenation of CO2 into methanol occurs via the following sequence: CO2* → HCOO* → HCOOH* → CH3O2* → CH2O* → CH3O*→CH3OH* where the symbol * indicates an adsorbed species.
TE D
A mechanistic study by Studt et al. [66], which compared CO2 to methanol pathways on different Cu-based catalysts, shed further light on the debate regarding the importance of differences in catalyst composition impacting on the resulting mechanism experienced by the
EP
CO2 molecules. In addition to the known effects of variation in reaction conditions (e.g. feed gas composition, temperature, and pressure), their calculations demonstrated that the
AC C
promoting effect of Zn in a Cu-ZnO containing catalyst was not only kinetically-based, but also stemmed from a deviation in the reaction mechanism due to the different sites at which the key reactions occur.
4. Available candidates and opportunities in catalyst materials selection Materials that have been studied for CO2 hydrogenation is listed in Table 1. One of the most common families of materials used for general catalysis is the metal family. For CO2 conversion, transition metals such as platinum [67, 68], copper [63, 65, 66, 69, 70], cobalt [71, 19
ACCEPTED MANUSCRIPT 72], ruthenium [56, 73, 74], rhodium [75-79], and nickel [53, 79-85] are amongst those found to be active for CO2 hydrogenation. Aside from being applied singly, the alloying of different metals to form a bimetallic catalyst has also been exploited to manipulate a catalyst’s electronic properties [86-91]. One of the primary aims of doing so is to affect product
RI PT
selectivity. For example, on combining platinum and cobalt a greater tendency toward CO production was reported [89] whereas varying the composition of a cobalt-iron bimetallic catalyst influenced the resulting alkane/alcohol selectivity [91]. More recently, a nickel-
SC
gallium catalyst [90] was developed and shown to be capable of higher selectivity toward
M AN U
methanol production.
Due to the effect of particle size on catalyst activity, nano-scaled metal catalyst size is often obtained and/or maintained by its dispersion upon a support material. Rather than acting simply as an inert host, such supports are often capable of exerting significant influence over the resulting surface chemistry of a catalyst. For the purpose of CO2 conversion, one of the
TE D
most useful traits of a support is its ability to encourage the adsorption and even the activation of the CO2 reactant. Such observations have been reported for conventionally
EP
popular oxide-type supports. While ZnO is one of the most studied oxide supports for (thermal) CO2 catalysis, due to its synergistic effects with copper catalysts [51] [92-94],
AC C
investigations on surface defect-rich oxides such as TiO2 [43, 95-97], CeO2 [86, 98, 99], Ga2O3 [100, 101], In2O3 [102-105], and their combinations [65, 106-108] have also been gaining momentum. Similar CO2 activation capabilities were found for carbide materials such as titanium carbide [109, 110] and molybdenum carbides [111-113], where CO2 conversions were also obtained despite the absence of metals. Nevertheless, when coupled with metals such as copper, nickel, or gold, a significant enhancement in activity (with varying results in product selectivity) have been observed [109]. Such synergism is often attributed to the interaction between the metal catalyst and the oxide or carbide surface via occurrences such 20
ACCEPTED MANUSCRIPT as creation of active centres at the metal-support interface [43, 108, 114], promoted CO2 adsorption upon the oxide or carbide surface [106, 115] and charge redistribution at the metal-support interface.
RI PT
Aside from the conversion aspects, there is also an opportunity to adapt solar energy absorbing functionalities to these active catalyst materials. For example, amongst the metals active for CO2 reduction, Cu, Ni, and Au nanoparticles exhibit plasmonic absorption that lies within the solar spectrum [30, 116-121]. The concept of plasmonic CO2 catalysis has, in fact,
SC
been demonstrated in the works of Navarrete et al. [122] and Wang et al. [123]; the former
M AN U
presenting a proof-of-concept Cu/ZnO-impregnated aerogel-containing microreactor for selfheating CO2 reduction and the latter detailing a study on the effectiveness of laser-induced heating on Au/ZnO for a similar application.
Aside from the dependence on the metal component, there also exist prospects for the non-
TE D
metallic (support) materials to contribute to the harvesting of solar energy. Transition metal oxides, such as those mentioned above, are semiconductors that are capable of light energy absorption due to their unique band structures. Therefore, as described by Baffou and
EP
Quidant [31], in addition to thermoplasmonic effects, the combination of an active, plasmonic metal catalyst with a semiconducting material can result in enhanced charge carrier
AC C
generation (due to thermoplasmonics or optical near-field enhancement) or hot electron injection into the semiconductor (Figure 5d-f). Such interactions upon light illumination have been evidenced in various works on both catalytic and mechanistic studies [124-130]. Additionally, it is not uncommon for metal/metal oxide-type materials to be used as selective absorber coatings for solar thermal collectors. The coupling of aluminium oxide with metals such as cobalt [131, 132], silver [133], platinum [134], molybdenum [135, 136], and most prevalently, nickel [137-141] has been previously reported. The popularity of the nickel-
21
ACCEPTED MANUSCRIPT alumina composite stems from its solar absorption properties – it possesses the ideal combination of high absorbance (for effective energy absorption) and low thermal emissivity (to minimise losses through thermal radiation) [138]. The possibility of utilising the spectrum-selective characteristics of such a combination of materials for solar thermal
RI PT
catalysis has been revealed to be attainable through the CAESAR (CAtalytically Enhanced Solar Absorption Receiver) project [142]. CAESAR was applied to the CO2 reforming of methane using a catalyst material (in the form of a catalyst foam), which acted as a receiver
SC
upon which the concentrated solar radiation was reflected to. Carbide materials also exhibit great potential for such an application. Carbides that have been tested for selective solar
M AN U
absorbance purposes include hafnium, tantalum, and titanium carbides [143-145]. Although studies specifically on carbides active for CO2 reduction (e.g. Ti- and Mo-carbides) are less common than that of alumina, the carbide family of refractory materials has the distinctive advantage of high thermal resistance [146]. This is an important attribute for fabricating a
TE D
catalyst/absorber foam with greater durability as the loss of structural integrity is frequently cited as the mode of failure for alumina-based catalyst foams [16, 17].
AC C
EP
Table 1. Catalyst materials reported for CO2 hydrogenation reactions and their respective performance. Catalyst Reaction Products Performance Ref conditions 1 wt% Pt/TiO2 [43] • Fixed bed • CO (~50 %) • 2.03 - 6.48 × 10−3mol s-1 flow gcat-1 (573 K) • Remaining • 573 – 873 K composition • TOF = 0.423 – 2.716 mol s1 not (573 K) explicitly mentioned • Fixed bed • CO – major • TOF = 0.44 – 2.716 s-1 (573 [95] flow product K) • 576 – 873 K • < 10 % CH4 production • Other products present but not explicitly mentioned
22
ACCEPTED MANUSCRIPT
Ni/Ni(Al)Ox
• • • • • •
NiO/SBA-15
2 wt% Pt/CeO2
Co/CeO2 promoted)
(K- • • •
Pt, Ni, Pd, Co on • CeO2
Fixed bed 673 – 1173 K Fixed bed 573 – 723 K Fixed bed In operando DRIFT-MS system Packed-bed 673 – 873 K Atmospheric pressure
• •
Ni /CZ**
•
423 – 673 K
EP • •
• •
Fixed bed 423 – 673 K
• • • Ce2O3-promoted • In2O3 • CeO2-promoted Ga- •
Packed bed 473 – 923 K
AC C • • •
Fixed bed Atmospheric pressure 673 – 753 K Fixed bed 423 – 673 K
Ni/Ce0.72Zr0.28O2
Ni-Rh/CZ**
In2O3
• 1.62 – 4.18x10-3 min-1 gcat-1
[86]
• •
CH4 CO
• 0 – 80 % conversion
[81]
• • • • • • • • • •
CH4 CO CH4 Trace CO Trace C2H6 CO CH4 CH4 CO Pure CO products
• 0 – 75 %
[82]
• Approx. 90 % CO2 conversion
[83]
• 0 – 50 % CO2 conversion • TOF = 0 – 12 min-1 • 80 – 100 % CO2 conversion
[147]
Fixed bed 523 – 773 K Fixed bed
• -
[148] [51]
• •
CO CH4
• 13 – 38 % CO2 conversion
[98]
• •
CO CH4
• 2.88 – 12.51x10-3 min-1 gcat-
[86]
• •
CO CH4
[44]
• • • • • •
CH44 CO C2H6 CH4 CO C2H6
• 0.5 – 1.1 mmol gcat-1 min-1 (673 K) • 5.8- 20 mmol gcat-1 min-1 (823 K) • 70 – 80 % CO2 conversion • 2.16 – 2.41 mol CO2 gNi-1 s• Best activity at 76 % CO2 conversion
[149]
•
CH4
• 0.71 – 2.07 mol CO2 gNi-1 s-
[85]
TE D
NixCe0.75Zr0.25−xO2
Batch reactor with FTIR attachment 673 – 823 K 1 atm
CO CH4
RI PT
• • • •
• •
SC
• •
Ni/ Al2O3
Batch reactor with FTIR attachment Fixed bed 473 – 773 K, atmospheric pressure Fixed bed 523 – 773 K 523 – 773 K 5 – 20 bar
M AN U
Pt, Ni, Pd, Co on γ- • alumina
• • • •
CH4 CO C2H6 CO
•
CO
•
CO
1
[84]
1
1
• 41.1 – 85.2% CO2 conversion • 78 % CO2 conversion • 2.37 mol CO2 gNi-1 s-1
[84]
• 6 – 35 % CO2 conversion
[102]
• 0.39 – 19.98 % CO2 conversion • 2.58 – 10.99 % CO2
[107] [106]
23
Pt, Pd, Ni, Fe, or Cu • on La-ZrO2 • •
La1−xSrxCoO3-δ
• •
La0.75Sr0.25Co(1− y)FeYO3 TiC
• • • • •
ZrC
•
NbC
•
TaC
• • •
Mo2C
• • •
WC
•
•
CO Other products not mentioned CO ( 96 % selectivity) CH4
•
CO
•
• 30 – 35 % CO2 conversion
[151]
• 18.5 – 40.6 % CO2 conversion
[152]
[153]
• 110.8 µmol min-1 g-1 average CO generation rate
CO
• 40 – 80 µmol min-1 g-1
[154]
•
CO
• 0.59 % CO2 conversion • TOF = 2.1 min-1
[155]
AC C
• •
• •
[150]
•
EP
• •
Electric-field assisted thermal reaction Fixed bed 423 K with 3 mA input current RWGS-CL system* 1123 K (during CO2 feed) RWGS-CL system* 823 K Fixed bed reactor 573 K Atmospheric pressure Fixed bed reactor 573 K Atmospheric pressure Fixed bed reactor 573 K Atmospheric pressure Fixed bed reactor 573 K Atmospheric pressure Fixed bed reactor 573 K Atmospheric pressure Fixed bed reactor
CO (97 % selectivity)
RI PT
• •
•
conversion • 37.5 % (for BaZr0.8Y0.16Zn0.04O3)
M AN U
FeO nanoparticles
523 – 773 K Fixed bed 873 K Atmospheric pressure Fixed bed 873 K
TE D
22O3 • BaZrO3 (Y, Zn, Ce- • doped) • •
SC
ACCEPTED MANUSCRIPT
•
CO
• 0.46 % CO2 conversion • TOF = 8.9 min-1
[155]
• •
CO CH4
• 2.09 % CO2 conversion • TOF = 61.1 min-1
[155]
•
CO
• 1.73 % CO2 conversion • TOF = 52.8 min-1
[155]
• • •
CH4 CO
• 4.67 % CO2 • TOF = 66.5 min-1
[155]
• •
CO CH4
• 3.30 % CO2 conversion • TOF = 59.1 min-1
[155]
24
ACCEPTED MANUSCRIPT
CO CH3OH C2H5OH CH3OCH3 CH4 C2H6 C3H8 and others CO (major product) CH4 CH3OH
Batch reactor • attached to UHV chamber • • 0.5 atm • CO2:4.5 atm H2 * RWGS-CL: Reverse water-gas shift chemical looping ** CZ: ceria-zirconia
• 4 – 31 % conversion
• 10 – 14x1015 produced molecules cm-2 s-1
[112]
[109]
M AN U
Au/TiC
•
• • • • • • •
RI PT
Cu, Ni, or Co on • Mo2C •
573 K Atmospheric pressure 473 – 575 K 2 MPa
SC
• •
5. Challenges in the execution of catalytic photo-thermal CO2 reduction An ideal catalyst possesses the following qualities: high activity, selectivity, and stability. To design a photo-thermal catalyst whereby heat generation is supplemented, partially if not
TE D
completely, from its solar absorption capabilities, additional considerations have to be taken into account for the material’s heat generation efficiencies.
EP
For example, the issue of material stability becomes even more substantial in the case of plasmonic catalysis as plasmonic absorption and the resulting effects are greatly affected by
AC C
both the size and shape of the nanoparticles [20, 21, 156]. It is also known that nanoparticles tend to succumb to events such as sintering and loss of nanostructured morphology upon extensive exposure to heat. Therefore, the control of such properties of a material becomes one of the key steps to ensure the practical applicability and robustness of the phenomenon. The design of photo-thermal materials often calls for a composite component to facilitate optimisation of the solar heat generation performance and catalytic conversion capabilities. In this case, the effect each component has on the other also must be taken into account. For
25
ACCEPTED MANUSCRIPT instance, in the case of Au/TiO2 whereby both the metal and support are capable of lightactivation, Tan et al. [157] used in situ spectroscopic studies to illustrate a difference in the preferred pathway (for ethanol oxidation) upon introducing a light source to a thermal reaction. In addition, on utilising light sources with different wavelength ranges (ultraviolet
RI PT
and visible), variations in the origins of the active electrons and their subsequent transfer routes were observed, which consequently altered the reaction outcome. The study highlighted the importance of understanding the pathways associated with photo-thermal CO2
SC
reactions to gain better insights into catalyst performance especially in terms of product
M AN U
selectivity.
Complementary to material design and fabrication, the experimental setup for photo-thermal catalysis requires careful consideration, as conventional systems used for gas-solid heterogeneous catalysis (e.g. enclosed plug-flow tube reactors) are often not suited to lightbased applications. Existing solar thermal reactors, such as the one designed by Palumbo et al.
TE D
(Figure 10), can be modified for photo-thermal catalyses under concentrated solar irradiation [158]. In their study, Meng et al. [159] retrofitted a water splitting rig for tandem hydrogen
EP
generation and photo-thermal CO2 reduction using a xenon arc lamp (20 mW/cm2) as the illumination source, Nonetheless, the main drawback of these designs is the lack of a
AC C
secondary heating system for deconvoluting photothermal heating and photo-induced charge transfer effects. To achieve this, both Westrich et al. and Tan et al. utilised external heating elements to heat up the catalyst bed from the base while exposing the top to continuous illumination [160, 161]. Alternatively, Upadhye et al. [162] adapted a commercially available reactor (Harrick Scientific, HVC-MRA-5), which was originally designed for in-situ spectroscopy analyses, to study the impact of visible light illumination on CO2 reduction by oxide-supported Au catalysts. In these studies, the secondary heating systems were capable of generating temperatures up to 400 – 600 oC, which allows for the study of non-illuminated 26
ACCEPTED MANUSCRIPT (i.e. dark) thermal catalysis, while thermal heating from the irradiation sources was
M AN U
SC
RI PT
minimised through the use of specific bandwidth sources and/or filters.
TE D
Figure 10 Potential reactor configurations for coupling and/or decoupling of heat and light in photo-thermal catalytic studies: (a) Palumbo et al. [158]; (b) Meng et al. [159]; (c) Westrich et al. (Adapted with permission from ref. [160]. Copyright (2011) American Chemical Society.); (d) Tan et al. (Adapted with permission from ref. [161]. Copyright (2016) American Chemical Society.); and (e) Upadhye et al. (Reproduced from ref. [162] with permission from The Royal Society of Chemistry.).
While the focus thus far has been placed on the reduction of CO2 to fuels, it must be kept in
EP
mind that hydrogen plays an equally important role in the reaction. This then brings us to the
AC C
issue of a suitable hydrogen source. Despite the increasing abundance of CO2, hydrogen is not as readily available and requires additional steps for its production. In terms of current approaches, the steam reforming of methane remains as the primary route for its large-scale production. However, despite its effectiveness, steam reforming is an energy intensive process often powered by fossil fuels. This will then further contribute to CO2 emissions and defeat the original intention of utilising hydrogen for CO2 recycling. Aside from natural gas, water is another appealing hydrogen source. In order to overcome the energy intensive water splitting process for effective mass hydrogen (and oxygen) production, 27
ACCEPTED MANUSCRIPT three potential solutions are frequently proposed: photocatalytic water splitting [163, 164], photoelectrochemical water splitting [165] and solar powered electrocatalytic water splitting [166]. Hydrogen can be produced by photocatalysis via the direct utilisation of solar energy. Reviews detailing the studies and applications of sustainable photocatalysts to achieve cost-
RI PT
effective photocatalytic water splitting are readily available [164, 167]. Recent developments in photocatalysis include the introduction of carbon-based nanostructures coupled with existing semiconductors or as standalone photocatalysts [163, 164, 168-170]. While
SC
exhibiting potential, separation of the hydrogen and oxygen gas products remains a major
M AN U
bottleneck for coupling this technology with photo-thermal CO2 recycling. Another potentially sustainable hydrogen generation technology is photoelectrochemical water splitting which is essentially a short-circuited solar-powered electrochemical cell. Photoelectrochemical water splitting utilises semiconductor-based electrodes, which are capable of splitting water into hydrogen and oxygen upon photoexcitation. Nevertheless, the
TE D
viability of this solution for large-scale application is often challenged by its low solar energy to electricity conversion. To date, electrocatalysis coupled with matured photovoltaic
EP
technology is thought to be the most feasible system for mass producing hydrogen as a feed for CO2 recycling. The conversion record is currently held at 40 % efficiency with the use of
AC C
a triple junction solar cell [171]. The feasibility of such solar-powered electrocatalytic water splitting has, in fact, been evidenced by Jia et al. [166], whose work demonstrated a solar-tohydrogen conversion capability of more than 30 %.
6. Conclusion and outlook Solar-powered fuel production that consumes CO2 as feedstock is an ideal concept that shows great potential for simultaneous CO2 mitigation and renewable fuel production. While the thermal catalytic reduction of CO2 has been relatively well-studied, the effects of coupling 28
ACCEPTED MANUSCRIPT the catalytic process with light requires further work and understanding in order to gain a better comprehension of the process chemistry and mechanisms involved. Although solar thermal technologies and plasmonic catalysis have mostly been discussed
RI PT
individually, they do not have to be mutually exclusive. Concurrent use of both technologies may, in fact, complement the shortcomings of each technique. For example, the utilisation of plasmonic materials in a solar thermal reactor may be used to reduce the amount of solar concentration required from the collector, thus reducing the severe requirements on the solar
M AN U
generation to achieve a sufficient temperature.
SC
reactor infrastructure while reducing the sole dependence on thermoplasmonic heat
To accomplish the realisation of photo-thermal CO2 catalysis, one of the key factors is successful integration of multiple disciplines of study: solar absorbing infrastructure design; materials development; the sourcing of renewable hydrogen; effective CO2 capture; and an
TE D
understanding of CO2 conversion pathways. Furthermore, surveys on influential factors, such as solar irradiation consistency, space availability, and the integration of synthetic fuel production with the existing distribution network and infrastructure will also be necessary to
AC C
locations.
EP
assess the suitability of photo-thermal CO2 reduction processes for different geographical
29
ACCEPTED MANUSCRIPT
References
7. 8. 9. 10. 11.
12.
13.
14.
15. 16. 17. 18.
19.
20.
RI PT
6.
SC
5.
M AN U
4.
TE D
3.
EP
2.
ExxonMobil, 2017 outlook for energy: A view to 2040. 2017, Exxon Mobil Corporation: Texas, USA. Böhmer, M., U. Langnickel, and M. Sanchez, Solar steam reforming of methane. Solar Energy Materials, 1991. 24(1-4): p. 441-448. Fujishima, A. and K. Honda, TiO2 photoelectrochemistry and photocatalysis. Nature, 1972. 238(5358): p. 37-38. Agrafiotis, C., et al., Solar thermal reforming of methane feedstocks for hydrogen and syngas production—A review. Renewable and Sustainable Energy Reviews, 2014. 29: p. 656-682. Kalogirou, S.A., Solar thermal collectors and applications. Progress in Energy and Combustion Science, 2004. 30(3): p. 231-295. Kodama, T., High-temperature solar chemistry for converting solar heat to chemical fuels. Progress in Energy and Combustion Science, 2003. 29(6): p. 567-597. Bogaerts, W.F. and C.M. Lampert, Materials for photothermal solar energy conversion. Journal of Materials Science, 1983. 18(10): p. 2847-2875. Cao, F., et al., A review of cermet-based spectrally selective solar absorbers. Energy & Environmental Science, 2014. 7(5): p. 1615-1627. Chueh, W.C., et al., High-flux solar-driven thermochemical dissociation of CO2 and H2O using nonstoichiometric ceria. Science, 2010. 330(6012): p. 1797-1801. Furler, P., et al., Solar thermochemical CO2 splitting utilizing a reticulated porous ceria redox system. Energy & Fuels, 2012. 26(11): p. 7051-7059. Le Gal, A., S. Abanades, and G. Flamant, CO2 and H2O splitting for thermochemical production of solar fuels using nonstoichiometric ceria and ceria/zirconia solid solutions. Energy & Fuels, 2011. 25(10): p. 4836-4845. Abanades, S. and M. Chambon, CO2 dissociation and upgrading from two-step solar thermochemical processes based on ZnO/Zn and SnO2/SnO redox pairs. Energy & Fuels, 2010. 24(12): p. 6667-6674. Galvez, M., et al., CO2 splitting via two-step solar thermochemical cycles with Zn/ZnO and FeO/Fe3O4 redox reactions: thermodynamic analysis. Energy & Fuels, 2008. 22(5): p. 3544-3550. Loutzenhiser, P.G. and A. Steinfeld, Solar syngas production from CO2 and H2O in a two-step thermochemical cycle via Zn/ZnO redox reactions: thermodynamic cycle analysis. International Journal of Hydrogen Energy, 2011. 36(19): p. 12141-12147. Dahl, J.K., et al., Dry reforming of methane using a solar-thermal aerosol flow reactor. Industrial & engineering chemistry research, 2004. 43(18): p. 5489-5495. Muir, J.F., et al., Solar reforming of methane in a direct absorption catalytic reactor on a parabolic dish: I—Test and analysis. Solar Energy, 1994. 52(6): p. 467-477. Wörner, A. and R. Tamme, CO2 reforming of methane in a solar driven volumetric receiver–reactor. Catalysis Today, 1998. 46(2): p. 165-174. Kelly, K.L., et al., The optical properties of metal nanoparticles: the influence of size, shape, and dielectric environment. Journal of Physical Chemistry B-Condensed Phase, 2003. 107(3): p. 668-677. Zeman, E.J. and G.C. Schatz, An accurate electromagnetic theory study of surface enhancement factors for Ag, Au, Cu, Li, Na, Al, Ga, in, Zn, and Cd. Journal of Physical Chemistry, 1987. 91(3): p. 634-643. Link, S. and M.A. El-Sayed, Size and temperature dependence of the plasmon absorption of colloidal gold nanoparticles. The Journal of Physical Chemistry B, 1999. 103(21): p. 4212-4217.
AC C
1.
30
ACCEPTED MANUSCRIPT
27.
28. 29.
30. 31. 32. 33.
34.
35. 36. 37. 38. 39.
40. 41.
RI PT
26.
SC
25.
M AN U
24.
TE D
23.
EP
22.
Noguez, C., Surface plasmons on metal nanoparticles: the influence of shape and physical environment. The Journal of Physical Chemistry C, 2007. 111(10): p. 38063819. Lee, K.-S. and M.A. El-Sayed, Gold and silver nanoparticles in sensing and imaging: sensitivity of plasmon response to size, shape, and metal composition. The Journal of Physical Chemistry B, 2006. 110(39): p. 19220-19225. Mock, J., et al., Shape effects in plasmon resonance of individual colloidal silver nanoparticles. The Journal of Chemical Physics, 2002. 116(15): p. 6755-6759. Hodak, J.H., et al., Laser-induced inter-diffusion in AuAg core− shell nanoparticles. The Journal of Physical Chemistry B, 2000. 104(49): p. 11708-11718. Papoff, F. and B. Hourahine, Geometrical Mie theory for resonances in nanoparticles of any shape. Optics Express, 2011. 19(22): p. 21432-21444. Blaber, M.G., M.D. Arnold, and M.J. Ford, A review of the optical properties of alloys and intermetallics for plasmonics. Journal of Physics: Condensed Matter, 2010. 22(14): p. 143201. Cao, L., et al., Plasmon-assisted local temperature control to pattern individual semiconductor nanowires and carbon nanotubes. Nano Letters, 2007. 7(11): p. 35233527. Gogna, P.K., Chopra, K. L. & Mullick, S. C Photothermal performance of selective black nickel coatings. International Journal of Energy Research, 1980. 4: p. 317-322. Scherer, A., Inal, O. T. & Pettit, R. B. , Modelling of degradation of nickel-pigmented aluminium oxide photothermal collector coatings. Journal of Materials Science, 1988. 23: p. 1934-1942. Chen, X., et al., Nanosecond photothermal effects in plasmonic nanostructures. ACS Nano, 2012. 6(3): p. 2550-2557. Baffou, G. and R. Quidant, Nanoplasmonics for chemistry. Chemical Society Reviews, 2014. 43(11): p. 3898-3907. Tas, G.M., H. J., Electron diffusion in metals studied by picosecond ultrasonics. Physical Review B, 1994: p. 15046-15054. Manda Xiao, R.J., Feng Wang, Caihong Fang, Jianfang Wang and Jimmy C. Yu, Plasmon-enhanced chemical reactions. Journal of Material Chemistry A, 2013. 1: p. 5790-5805. Lo, C.-C., et al., Photoreduction of carbon dioxide with H2 and H2O over TiO2 and ZrO2 in a circulated photocatalytic reactor. Solar Energy Materials and Solar Cells, 2007. 91(19): p. 1765-1774. Zhang, X., et al., Plasmonic photocatalysis. Reports on Progress in Physics, 2013. 76(4): p. 046401. Teramura, K., et al., Photocatalytic conversion of CO2 in water over layered double hydroxides. Angewandte Chemie, 2012. 124(32): p. 8132-8135. Sanchez, A., et al., When gold is not noble: Nanoscale gold catalysts. The Journal of Physical Chemistry A, 1999. 103(48): p. 9573-9578. Anikeev, V.I., et al., Catalytic thermochemical reactor/receiver for solar reforming of natural gas: Design and performance. Solar Energy, 1998. 63(2): p. 97-104. Kok, E., et al., The effect of support and synthesis method on the methanation activity of alumina-supported cobalt–ruthenium–lanthana catalysts. Catalysis Today, 2011. 178(1): p. 79-84. Trimm, D., Coke formation and minimisation during steam reforming reactions. Catalysis Today, 1997. 37(3): p. 233-238. Cargnello, M., et al., Control of metal nanocrystal size reveals metal-support interface role for ceria catalysts. Science, 2013. 341(6147): p. 771-773.
AC C
21.
31
ACCEPTED MANUSCRIPT
48.
49. 50. 51.
52.
53.
54.
55.
56. 57.
58.
59. 60.
RI PT
47.
SC
46.
M AN U
45.
TE D
44.
EP
43.
Linic, S., Aslam, U., Boerigter, C. & Morabito, M. . Photochemical transformations on plasmonic metal nanoparticles. Nature Materials, 2015. 14: p. 567-576. Kim, S.S., H.H. Lee, and S.C. Hong, The effect of the morphological characteristics of TiO2 supports on the reverse water–gas shift reaction over Pt/TiO2 catalysts. Applied Catalysis B: Environmental, 2012. 119: p. 100-108. Zonetti, P.C., et al., The NixCe0.75Zr0.25−xO2 solid solution and the RWGS. Applied Catalysis A: General, 2014. 475: p. 48-54. Porosoff, M.D., B. Yan, and J.G. Chen, Catalytic reduction of CO2 by H2 for synthesis of CO, methanol and hydrocarbons: challenges and opportunities. Energy & Environmental Science, 2016. 9(1): p. 62-73. Fujita, S.-I., M. Usui, and N. Takezawa, Mechanism of the reverse water gas shift reaction over Cu/ZnO catalyst. Journal of Catalysis, 1992. 134(1): p. 220-225. Ginés, M., A. Marchi, and C. Apesteguia, Kinetic study of the reverse water-gas shift reaction over CuO/ZnO/Al2O3 catalysts. Applied Catalysis A: General, 1997. 154(12): p. 155-171. Chen, C.-S., W.-H. Cheng, and S.-S. Lin, Mechanism of CO formation in reverse water–gas shift reaction over Cu/Al2O3 catalyst. Catalysis Letters, 2000. 68(1): p. 4548. Chen, C.S., J.H. Wu, and T.W. Lai, Carbon dioxide hydrogenation on Cu nanoparticles. The Journal of Physical Chemistry C, 2010. 114(35): p. 15021-15028. Kung, H.H., Transition metal oxides : Surface chemistry and catalysis. 1989, Elsevier: Amsterdam, The Netherlands. Goguet, A., et al., Spectrokinetic investigation of reverse water-gas-shift reaction intermediates over a Pt/CeO2 catalyst. The Journal of Physical Chemistry B, 2004. 108(52): p. 20240-20246. Davis, W. and M. Martín, Optimal year-round operation for methane production from CO2 and water using wind and/or solar energy. Journal of Cleaner Production, 2014. 80: p. 252-261. Aldana, P.U., et al., Catalytic CO2 valorization into CH4 on Ni-based ceria-zirconia. Reaction mechanism by operando IR spectroscopy. Catalysis Today, 2013. 215: p. 201-207. Peebles, D., D. Goodman, and J. White, Methanation of carbon dioxide on nickel (100) and the effects of surface modifiers. The Journal of Physical Chemistry, 1983. 87(22): p. 4378-4387. Marwood, M., R. Doepper, and A. Renken, In-situ surface and gas phase analysis for kinetic studies under transient conditions The catalytic hydrogenation of CO2. Applied Catalysis A: General, 1997. 151(1): p. 223-246. Prairie, M.R., et al., A fourier transform infrared spectroscopic study of CO2 methanation on supported ruthenium. Journal of Catalysis, 1991. 129(1): p. 130-144. Jacquemin, M., A. Beuls, and P. Ruiz, Catalytic production of methane from CO2 and H2 at low temperature: Insight on the reaction mechanism. Catalysis Today, 2010. 157(1): p. 462-466. Vesselli, E., et al., Hydrogen-assisted transformation of CO2 on nickel: the role of formate and carbon monoxide. The Journal of Physical Chemistry Letters, 2009. 1(1): p. 402-406. Ren, J., et al., Insights into the mechanisms of CO2 methanation on Ni (111) surfaces by density functional theory. Applied Surface Science, 2015. 351: p. 504-516. Olah, G.A., G.S. Prakash, and A. Goeppert, Anthropogenic chemical carbon cycle for a sustainable future. Journal of the American Chemical Society, 2011. 133(33): p. 12881-12898.
AC C
42.
32
ACCEPTED MANUSCRIPT
67.
68. 69. 70. 71.
72.
73.
74.
75. 76. 77. 78.
79. 80.
RI PT
66.
SC
65.
M AN U
64.
TE D
63.
EP
62.
Ortelli, E., J. Wambach, and A. Wokaun, Methanol synthesis reactions over a CuZr based catalyst investigated using periodic variations of reactant concentrations. Applied Catalysis A: General, 2001. 216(1): p. 227-241. Jadhav, S.G., et al., Catalytic carbon dioxide hydrogenation to methanol: A review of recent studies. Chemical Engineering Research and Design, 2014. 92(11): p. 25572567. Grabow, L. and M. Mavrikakis, Mechanism of methanol synthesis on Cu through CO2 and CO hydrogenation. ACS Catalysis, 2011. 1(4): p. 365-384. Hansen, J.B. and P.E.H. Nielsen, Handbook of heterogeneous catalysis, ed. G. Ertl, et al. 2008: Wiley. Graciani, J., et al., Highly active copper-ceria and copper-ceria-titania catalysts for methanol synthesis from CO2. Science, 2014. 345(6196): p. 546-550. Studt, F., et al., The mechanism of CO and CO2 hydrogenation to methanol over Cu based catalysts. ChemCatChem, 2015. 7(7): p. 1105-1111. Dietz, L., S. Piccinin, and M. Maestri, Mechanistic insights into CO2 activation via reverse water–gas shift on metal surfaces. The Journal of Physical Chemistry C, 2015. 119(9): p. 4959-4966. Yu, K.-P., et al., Pt/titania-nanotube: A potential catalyst for CO2 adsorption and hydrogenation. Applied Catalysis B: Environmental, 2008. 84(1): p. 112-118. Ahouari, H., et al., Methanol synthesis from CO2 hydrogenation over copper based catalysts. Reaction Kinetics, Mechanisms and Catalysis, 2013. 110(1): p. 131-145. Behrens, M., Heterogeneous catalysis of CO2 conversion to methanol on copper surfaces. Angewandte Chemie International Edition, 2014. 53(45): p. 12022-12024. Iablokov, V., et al., Size-controlled model co nanoparticle catalysts for CO2 hydrogenation: synthesis, characterization, and catalytic reactions. Nano Letters, 2012. 12(6): p. 3091-3096. Srisawad, N., et al., CO2 hydrogenation over Co/Al2O3 catalysts prepared via a solidstate reaction of fine gibbsite and cobalt precursors. Reaction Kinetics, Mechanisms and Catalysis, 2012. 107(1): p. 179-188. Kwak, J.H., L. Kovarik, and J.n. Szanyi, CO2 reduction on supported Ru/Al2O3 catalysts: cluster size dependence of product selectivity. ACS Catalysis, 2013. 3(11): p. 2449-2455. Roldán, L., Y. Marco, and E. García Bordejé, Function of the support and metal loading on catalytic carbon dioxide reduction using ruthenium nanoparticles supported on carbon nanofibers. ChemCatChem, 2015. 7(8): p. 1347-1356. Solymosi, F., A. Erdöhelyi, and T. Bansagi, Methanation of CO2 on supported rhodium catalyst. Journal of Catalysis, 1981. 68(2): p. 371-382. Beuls, A., et al., Methanation of CO2: Further insight into the mechanism over Rh/γAl2O3 catalyst. Applied Catalysis B: Environmental, 2012. 113: p. 2-10. Karelovic, A. and P. Ruiz, Mechanistic study of low temperature CO2 methanation over Rh/TiO2 catalysts. Journal of Catalysis, 2013. 301: p. 141-153. Karelovic, A. and P. Ruiz, Improving the hydrogenation function of Pd/γ-Al2O3 catalyst by Rh/γ-Al2O3 addition in CO2 methanation at low temperature. ACS Catalysis, 2013. 3(12): p. 2799-2812. Hubble, R., J. Lim, and J. Dennis, Kinetic studies of CO2 methanation over a Ni/γAl2O3 catalyst. Faraday Discussions, 2016. 192: p. 529-544. da Silva, D.C., et al., The Ni/ZrO2 catalyst and the methanation of CO and CO2. International Journal of Hydrogen Energy, 2012. 37(11): p. 8923-8928.
AC C
61.
33
ACCEPTED MANUSCRIPT
86.
87.
88.
89.
90. 91. 92.
93. 94.
95.
96.
97.
RI PT
SC
85.
M AN U
84.
TE D
83.
EP
82.
Rahmani, S., M. Rezaei, and F. Meshkani, Preparation of highly active nickel catalysts supported on mesoporous nanocrystalline γ-Al2O3 for CO2 methanation. Journal of Industrial and Engineering Chemistry, 2014. 20(4): p. 1346-1352. Garbarino, G., et al., A study of the methanation of carbon dioxide on Ni/Al2O3 catalysts at atmospheric pressure. International Journal of Hydrogen Energy, 2014. 39(22): p. 11557-11565. Abelló, S., C. Berrueco, and D. Montané, High-loaded nickel–alumina catalyst for direct CO2 hydrogenation into synthetic natural gas (SNG). Fuel, 2013. 113: p. 598609. Ocampo, F., et al., Effect of Ce/Zr composition and noble metal promotion on nickel based CexZr1− xO2 catalysts for carbon dioxide methanation. Applied Catalysis A: General, 2011. 392(1): p. 36-44. Ocampo, F., B. Louis, and A.-C. Roger, Methanation of carbon dioxide over nickelbased Ce0.72Zr0.28O2 mixed oxide catalysts prepared by sol–gel method. Applied Catalysis A: General, 2009. 369(1): p. 90-96. Porosoff, M.D. and J.G. Chen, Trends in the catalytic reduction of CO2 by hydrogen over supported monometallic and bimetallic catalysts. Journal of Catalysis, 2013. 301: p. 30-37. Shin, H.H., et al., Cobalt catalysts decorated with platinum atoms supported on barium zirconate provide enhanced activity and selectivity for CO2 methanation. ACS Catalysis, 2016. 6(5): p. 2811-2818. Zhu, Y., et al., Catalytic conversion of carbon dioxide to methane on ruthenium– cobalt bimetallic nanocatalysts and correlation between surface chemistry of catalysts under reaction conditions and catalytic performances. ACS Catalysis, 2012. 2(11): p. 2403-2408. Alayoglu, S., et al., CO2 hydrogenation studies on Co and CoPt bimetallic nanoparticles under reaction conditions using TEM, XPS and NEXAFS. Topics in catalysis, 2011. 54(13-15): p. 778. Studt, F., et al., Discovery of a Ni-Ga catalyst for carbon dioxide reduction to methanol. Nature chemistry, 2014. 6(4): p. 320-324. Gnanamani, M.K., et al., Hydrogenation of carbon dioxide over Co–Fe bimetallic catalysts. ACS Catalysis, 2016. 6(2): p. 913-927. Burch, R., S.E. Golunski, and M.S. Spencer, The role of copper and zinc oxide in methanol synthesis catalysts. Journal of the Chemical Society, Faraday Transactions, 1990. 86(15): p. 2683-2691. Behrens, M., et al., The active site of methanol synthesis over Cu/ZnO/Al2O3 industrial catalysts. Science, 2012. 336(6083): p. 893-897. Yoshihara, J. and C.T. Campbell, Methanol synthesis and reverse water–gas shift kinetics over Cu (110) model catalysts: Structural sensitivity. Journal of Catalysis, 1996. 161(2): p. 776-782. Kim, S.S., K.H. Park, and S.C. Hong, A study of the selectivity of the reverse water– gas-shift reaction over Pt/TiO2 catalysts. Fuel Processing Technology, 2013. 108: p. 47-54. Tada, S., et al., Ni/CeO2 catalysts with high CO2 methanation activity and high CH4 selectivity at low temperatures. International Journal of Hydrogen Energy, 2012. 37(7): p. 5527-5531. Abe, T., et al., CO2 methanation property of Ru nanoparticle-loaded TiO2 prepared by a polygonal barrel-sputtering method. Energy & Environmental Science, 2009. 2(3): p. 315-321.
AC C
81.
34
ACCEPTED MANUSCRIPT Wang, L., et al., K-promoted Co–CeO2 catalyst for the reverse water–gas shift reaction. Chemistry Letters, 2013. 42(7): p. 682-683. 99. Trovarelli, A., et al., CO2 methanation under transient and steady-state conditions over Rh/CeO2 and CeO2-promoted Rh/SiO2: The role of surface and bulk ceria. Journal of Catalysis, 1995. 151(1): p. 111-124. 100. Pan, Y.-x., et al., Hydrogen adsorption on Ga2O3 surface: A combined experimental and computational study. The Journal of Physical Chemistry C, 2011. 115(20): p. 10140-10146. 101. Qu, J., S.C.E. Tsang, and X.-Q. Gong, A DFT study on surface dependence of βGa2O3 for CO2 hydrogenation to CH3OH. Journal of Molecular Modeling, 2014. 20(12): p. 2543. 102. Sun, Q., et al., In2O3 as a promising catalyst for CO2 utilization: A case study with reverse water gas shift over In2O3. Greenhouse Gases: Science and Technology, 2014. 4(1): p. 140-144. 103. Ye, J., C. Liu, and Q. Ge, DFT study of CO2 adsorption and hydrogenation on the In2O3 surface. The Journal of Physical Chemistry C, 2012. 116(14): p. 7817-7825. 104. Ye, J., et al., Active oxygen vacancy site for methanol synthesis from CO2 hydrogenation on In2O3 (110): A DFT study. ACS Catalysis, 2013. 3(6): p. 1296-1306. 105. Sun, K., et al., Hydrogenation of CO 2 to methanol over In 2 O 3 catalyst. Journal of CO2 Utilization, 2015. 12: p. 1-6. 106. Zhao, B., Y.-x. Pan, and C.-j. Liu, The promotion effect of CeO2 on CO2 adsorption and hydrogenation over Ga2O3. Catalysis Today, 2012. 194(1): p. 60-64. 107. Wang, W., et al., Reverse water gas shift over In2 O3–CeO2 catalysts. Catalysis Today, 2016. 259: p. 402-408. 108. Yang, X., et al., Low Pressure CO2 hydrogenation to methanol over gold nanoparticles activated on a CeOx/TiO2 interface. Journal of the American Chemical Society, 2015. 137(32): p. 10104-10107. 109. Rodriguez, J.A., et al., CO2 hydrogenation on Au/TiC, Cu/TiC, and Ni/TiC catalysts: Production of CO, methanol, and methane. Journal of Catalysis, 2013. 307: p. 162169. 110. Vidal, A.B., et al., CO2 activation and methanol synthesis on novel Au/TiC and Cu/TiC catalysts. The Journal of Physical Chemistry Letters, 2012. 3(16): p. 22752280. 111. Porosoff, M.D., et al., Molybdenum carbide as alternative catalysts to precious metals for highly selective reduction of CO2 to CO. Angewandte Chemie International Edition, 2014. 53(26): p. 6705-6709. 112. Xu, W., et al., The carburization of transition metal molybdates (MxMoO4, M= Cu, Ni or Co) and the generation of highly active metal/carbide catalysts for CO2 hydrogenation. Catalysis Letters, 2015. 145(7): p. 1365-1373. 113. Posada-Pérez, S., et al., The conversion of CO2 to methanol on orthorhombic β-Mo2C and Cu/β-Mo2C catalysts: mechanism for admetal induced change in the selectivity and activity. Catalysis Science & Technology, 2016. 114. Lunkenbein, T., et al., Formation of a ZnO overlayer in industrial Cu/ZnO/Al2O3 catalysts induced by strong metal–support interactions. Angewandte Chemie, 2015. 127(15): p. 4627-4631. 115. Yang, Y., et al., Fundamental studies of methanol synthesis from CO2 hydrogenation on Cu (111), Cu clusters, and Cu/ZnO (0001). Physical Chemistry Chemical Physics, 2010. 12(33): p. 9909-9917.
AC C
EP
TE D
M AN U
SC
RI PT
98.
35
ACCEPTED MANUSCRIPT
121. 122.
123. 124.
125.
126.
127.
128.
129. 130.
131. 132.
RI PT
SC
120.
M AN U
119.
TE D
118.
EP
117.
Bracey, C.L., P.R. Ellis, and G.J. Hutchings, Application of copper–gold alloys in catalysis: current status and future perspectives. Chemical Society Reviews, 2009. 38(8): p. 2231-2243. Chan, G.H., et al., Plasmonic properties of copper nanoparticles fabricated by nanosphere lithography. Nano Letters, 2007. 7(7): p. 1947-1952. Sarina, S., et al., Enhancing catalytic performance of palladium in gold and palladium alloy nanoparticles for organic synthesis reactions through visible light irradiation at ambient temperatures. Journal of the American Chemical Society, 2013. 135(15): p. 5793-5801. Sugano, Y., et al., Supported Au–Cu bimetallic alloy nanoparticles: An aerobic oxidation catalyst with regenerable activity by visible light irradiation. Angewandte Chemie International Edition, 2013. 52(20): p. 5295-5299. Kowalska, E., et al., Visible-light-induced photocatalysis through surface plasmon excitation of gold on titania surfaces. Physical Chemistry Chemical Physics, 2010. 12(10): p. 2344-2355. Pirzadeh, Z., et al., Plasmon–interband coupling in nickel nanoantennas. ACS Photonics, 2014. 1(3): p. 158-162. Navarrete, A., et al., Novel windows for “solar commodities”: A device for CO2 reduction using plasmonic catalyst activation. Faraday Discussions, 2015. 183: p. 249-259. Wang, C., et al., Visible light plasmonic heating of Au–ZnO for the catalytic reduction of CO2. Nanoscale, 2013. 5(15): p. 6968-6974. Christopher, P., D.B. Ingram, and S. Linic, Enhancing photochemical activity of semiconductor nanoparticles with optically active Ag nanostructures: photochemistry mediated by Ag surface plasmons. The Journal of Physical Chemistry C, 2010. 114(19): p. 9173-9177. Gomes Silva, C.u., et al., Influence of excitation wavelength (UV or visible light) on the photocatalytic activity of titania containing gold nanoparticles for the generation of hydrogen or oxygen from water. Journal of the American Chemical Society, 2010. 133(3): p. 595-602. Awazu, K., et al., A plasmonic photocatalyst consisting of silver nanoparticles embedded in titanium dioxide. Journal of the American Chemical Society, 2008. 130(5): p. 1676-1680. Tian, Y. and T. Tatsuma, Mechanisms and applications of plasmon-induced charge separation at TiO2 films loaded with gold nanoparticles. Journal of the American Chemical Society, 2005. 127(20): p. 7632-7637. Tian, Y. and T. Tatsuma, Plasmon-induced photoelectrochemistry at metal nanoparticles supported on nanoporous TiO2. Chemical Communications, 2004(16): p. 1810-1811. Denzler, D.N., et al., Electronic excitation and dynamic promotion of a surface reaction. Physical review letters, 2003. 91(22): p. 226102. Mizeikis, V., E. Kowalska, and S. Juodkazis, Resonant localization, enhancement, and polarization of optical fields in nano-scale interface regions for photo-catalytic applications. Journal of Nanoscience and Nanotechnology, 2011. 11(4): p. 2814-2822. Niklasson, G. and C. Granqvist, Solar absorptance and thermal emittance of coevaporated Co-Al2O3 cermet films. Solar Energy Materials, 1983. 7(4): p. 501-510. Niklasson, G.A. and C.G. Granqvist, Optical properties and solar selectivity of coevaporated Co Al2O3 composite films. Journal of Applied Physics, 1984. 55(9): p. 3382-3410.
AC C
116.
36
ACCEPTED MANUSCRIPT
139. 140.
141.
142.
143. 144. 145. 146. 147. 148. 149.
150.
151.
152.
RI PT
138.
SC
137.
M AN U
136.
TE D
135.
EP
134.
Barshilia, H.C., et al., Structure and optical properties of Ag–Al2 O3 nanocermet solar selective coatings prepared using unbalanced magnetron sputtering. Solar Energy Materials and Solar Cells, 2011. 95(7): p. 1707-1715. Nuru, Z., et al., Thermal stability of multilayered Pt-Al2O 3 nanocoatings for high temperature CSP systems. Vacuum, 2015. 120: p. 115-120. Teixeira, V., et al., Spectrally selective composite coatings of Cr–Cr2O3 and Mo– Al2O3 for solar energy applications. Thin Solid Films, 2001. 392(2): p. 320-326. Zhang, Q.-C., Y. Yin, and D.R. Mills, High efficiency Mo-Al2O3 cermet selective surfaces for high-temperature application. Solar Energy Materials and Solar Cells, 1996. 40(1): p. 43-53. Zemanova, M., et al., Nickel electrolytic colouring of anodic alumina for selective solar absorbing films. Renewable Energy, 2008. 33(10): p. 2303-2310. Li, Z., J. Zhao, and L. Ren, Aqueous solution-chemical derived Ni–Al2O3 solar selective absorbing coatings. Solar Energy Materials and Solar Cells, 2012. 105: p. 90-95. Boström, T., E. Wäckelgård, and G. Westin, Solution-chemical derived nickel– alumina coatings for thermal solar absorbers. Solar Energy, 2003. 74(6): p. 497-503. Wang, X., et al., High-performance solution-processed plasmonic Ni nanochainAl2O3 selective solar thermal absorbers. Applied Physics Letters, 2012. 101(20): p. 203109. Galione, P., et al., Origin of solar thermal selectivity and interference effects in nickel–alumina nanostructured films. Surface and Coatings Technology, 2010. 204(14): p. 2197-2201. Buck, R., J.F. Muir, and R.E. Hogan, Carbon dioxide reforming of methane in a solar volumetric receiver/reactor: the CAESAR project. Solar Energy Materials, 1991. 24(1-4): p. 449-463. Sani, E., et al., Hafnium and tantalum carbides for high temperature solar receivers. Journal of Renewable and Sustainable Energy, 2011. 3(6): p. 063107. Lampert, C.M., Coatings for enhanced photothermal energy collection I. Selective absorbers. Solar Energy Materials, 1979. 1(5-6): p. 319-341. Harding, G., Sputtered metal carbide solar-selective absorbing surfaces. Journal of Vacuum Science and Technology, 1976. 13(5): p. 1070-1072. Toth, L.E., Transition metal carbides and nitrides. Refractory materials, ed. J.L. Margrave. Vol. 7. 1971, New York: Academic Press. Lu, B. and K. Kawamoto, Direct synthesis of highly loaded and well-dispersed NiO/SBA-15 for producer gas conversion. RSC Advances, 2012. 2(17): p. 6800-6805. Lu, B. and K. Kawamoto, Preparation of the highly loaded and well-dispersed NiO/SBA-15 for methanation of producer gas. Fuel, 2013. 103: p. 699-704. Cai, W., Q. Zhong, and Y. Zhao, Fractional-hydrolysis-driven formation of nonuniform dopant concentration catalyst nanoparticles of Ni/CexZr1− xO2 and its catalysis in methanation of CO2. Catalysis Communications, 2013. 39: p. 30-34. Kim, D.H., et al., Dopant effect of barium zirconate-based perovskite-type catalysts for the intermediate-temperature reverse water gas shift reaction. ACS Catalysis, 2014. 4(9): p. 3117-3122. Kim, D.H., et al., Reverse water gas shift reaction catalyzed by Fe nanoparticles with high catalytic activity and stability. Journal of Industrial and Engineering Chemistry, 2015. 23: p. 67-71. Oshima, K., et al., Low temperature catalytic reverse water gas shift reaction assisted by an electric field. Catalysis Today, 2014. 232: p. 27-32.
AC C
133.
37
ACCEPTED MANUSCRIPT
158.
159.
160.
161.
162.
163.
164. 165.
166. 167. 168. 169.
RI PT
SC
157.
M AN U
156.
TE D
155.
EP
154.
Daza, Y.A., et al., Carbon dioxide conversion by reverse water–gas shift chemical looping on perovskite-type oxides. Industrial & Engineering Chemistry Research, 2014. 53(14): p. 5828-5837. Daza, Y.A., et al., Isothermal reverse water gas shift chemical looping on La0.75Sr0.25Co(1− Y)FeYO3 perovskite-type oxides. Catalysis Today, 2015. 258: p. 691698. Porosoff, M.D., et al., Identifying trends and descriptors for selective CO2 conversion to CO over transition metal carbides. Chemical Communications, 2015. 51(32): p. 6988-6991. Jain, P.K., et al., Calculated absorption and scattering properties of gold nanoparticles of different size, shape, and composition: applications in biological imaging and biomedicine. Journal of Physical Chemistry B, 2006. 110(14): p. 7238. Tan, T.H., et al., C–C cleavage by Au/TiO2 during ethanol oxidation: understanding bandgap photoexcitation and plasmonically mediated charge transfer via quantitative in situ DRIFTS. ACS Catalysis, 2016. 6(12): p. 8021-8029. Palumbo, R., Keunecke, M., Möller, S. & Steinfeld, A, Reflections on the design of solar thermal chemical reactors: thoughts in transformation. Energy, 2004. 29: p. 727-744. Xianguang Meng, T.W., Lequan Liu, Shuxin Ouyang, Peng Li, Huilin Hu, Tetsuya Kako, Hideo Iwai, Akihiro Tanaka, and Jinhua Ye, Photothermal conversion of CO2 into CH4 with H2 over Group VIII nanocatalysts: An alternative approach for solar fuel production. Angewandte Chemie, 2014: p. 11662-11666. Westrich, T.A., Dahlberg, K. A., Kaviany, M. & Schwank, J. W. H. J. Phys. Chem. C 115, 16537–16543 (2011). High-temperature photocatalytic ethylene oxidation over TiO2. Journal of Physical Chemistry C, 2011. 115: p. 16537-16543. Tan, T.H., Scott, J. Ng, Y. H., Taylor, R. A., Aguey-Zinsou, K.-F., Amal, R., Understanding plasmon and band gap photoexcitation effects on the thermal-catalytic oxidation of ethanol by TiO2-supported gold. ACS Catalysis, 2016. 6: p. 1870-1879. Aniruddha A. Upadhye, I.R., Xu Zeng, Hyung Ju Kim, Isabel Tejedor, Marc A. Anderson, James A. Dumesic and George W. Huber, Plasmon-enhanced reverse water gas shift reaction over oxide supported Au catalysts. Catalysis Science & Technology, 2015. 5: p. 2590-2601. Ng, Y.H., et al., Reducing graphene oxide on a visible-light BiVO4 photocatalyst for an enhanced photoelectrochemical water splitting. The Journal of Physical Chemistry Letters, 2010. 1(17): p. 2607-2612. Kudo, A. and Y. Miseki, Heterogeneous photocatalyst materials for water splitting. Chemical Society Reviews, 2009. 38(1): p. 253-278. Hisatomi, T., J. Kubota, and K. Domen, Recent advances in semiconductors for photocatalytic and photoelectrochemical water splitting. Chemical Society Reviews, 2014. 43(22): p. 7520-7535. Jia, J., et al., Solar water splitting by photovoltaic-electrolysis with a solar-tohydrogen efficiency over 30%. Nature Communications, 2016. 7. Ran, J., et al., Earth-abundant cocatalysts for semiconductor-based photocatalytic water splitting. Chemical Society Reviews, 2014. 43(22): p. 7787-7812. Ran, J., et al., Ti3C2 MXene co-catalyst on metal sulfide photo-absorbers for enhanced visible-light photocatalytic hydrogen production. Nature Communications, 2017. 8. Rahman, M.Z., et al., 2D phosphorene as a water splitting photocatalyst: Fundamentals to applications. Energy Environ. Sci., 2016. 9(3): p. 709-728.
AC C
153.
38
ACCEPTED MANUSCRIPT Ran, J., et al., Porous P-doped graphitic carbon nitride nanosheets for synergistically enhanced visible-light photocatalytic H2 production. Energy & Environmental Science, 2015. 8(12): p. 3708-3717. 171. Green, M.A., et al., Solar cell efficiency tables (Version 45). Progress in Photovoltaics: Research and Applications, 2015. 23(1): p. 1-9.
AC C
EP
TE D
M AN U
SC
RI PT
170.
39