Next Article in Journal
Environmental-Friendly Synthesis of Alkyl Carbamates from Urea and Alcohols with Silica Gel Supported Catalysts
Next Article in Special Issue
Metal Micro-Monoliths for the Kinetic Study and the Intensification of the Water Gas Shift Reaction
Previous Article in Journal
Identification of Key Amino Acid Residues Determining Product Specificity of 2,3-Oxidosqualene Cyclase in Siraitia grosvenorii
Previous Article in Special Issue
The Influence of the Washcoat Deposition Process on High Pore Density Open Cell Foams Activation for CO Catalytic Combustion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Policies and Motivations for the CO2 Valorization through the Sabatier Reaction Using Structured Catalysts. A Review of the Most Recent Advances

Departamento de Química Inorgánica, and Instituto de Ciencia de Materiales de Sevilla (Centro Mixto Universidad de Sevilla-CSIC), Av. Américo Vespucio 49, 41092 Seville, Spain
*
Authors to whom correspondence should be addressed.
Catalysts 2018, 8(12), 578; https://doi.org/10.3390/catal8120578
Submission received: 10 October 2018 / Revised: 8 November 2018 / Accepted: 16 November 2018 / Published: 22 November 2018
(This article belongs to the Special Issue Structured Catalysts for Catalytic Processes Intensification)

Abstract

:
The current scenario where the effects of global warming are more and more evident, has motivated different initiatives for facing this, such as the creation of global policies with a clear environmental guideline. Within these policies, the control of Greenhouse Gase (GHG) emissions has been defined as mandatory, but for carrying out this, a smart strategy is proposed. This is the application of a circular economy model, which seeks to minimize the generation of waste and maximize the efficient use of resources. From this point of view, CO2 recycling is an alternative to reduce emissions to the atmosphere, and we need to look for new business models which valorization this compound which now must be considered as a renewable carbon source. This has renewed the interest in known processes for the chemical transformation of CO2 but that have not been applied at industrial level because they do not offer evident profitability. For example, the methane produced in the Sabatier reaction has a great potential for application, but this depends on the existence of a sustainable supply of hydrogen and a greater efficiency during the process that allows maximizing energy efficiency and thermal control to maximize the methane yield. Regarding energy efficiency and thermal control of the process, the use of structured reactors is an appropriate strategy. The evolution of new technologies, such as 3D printing, and the consolidation of knowledge in the structing of catalysts has enabled the use of these reactors to develop a wide range of possibilities in the field. In this sense, the present review presents a brief description of the main policies that have motivated the transition to a circular economy model and within this, to CO2 recycling. This allows understanding, why efforts are being focused on the development of different reactions for CO2 valorization. Special attention to the case of the Sabatier reaction and in the application of structured reactors for such process is paid.

1. Introduction

Nowadays, we are facing a scenario in which the effects of global warming are advancing rapidly, and the search for solutions goes beyond a simple technological challenge [1]. More integrated solutions are needed that connect different areas and, thus, generate a change in the model of life that allows guaranteeing the future of the planet and the permanence of the human beings in a sustainable way. One of the first steps to launch the search for solutions to stop global warming is the approach of laws, agreements, and treaties at the global, regional, and local level, which recognize the problem, define the main topics and propose a strategy of action that incorporates the different sectors, including scientific and technological.
In this sense, the reduction of greenhouse gas (GHG) emissions is one of the objectives in the new strategies to curb climate change. However, such a challenge implies the modification of the current productive process and a deep re-evaluation of our lifestyle, which has social, economic, cultural, and political implications. This is why the circular economy model has been proposed as an alternative that seeks to make our relationship with the planet and its resources sustainable and guarantee a future for the new generations [2].
These new approaches have led, for example, to viewing CO2 as no longer waste but rather a raw material with which you can obtain value-added products among which is a substitute for natural gas [3]. This explains the renewed interest in the reactions of the transformation of CO2 that, until now, have only been involved in industrial processes in the treatment of relatively small amounts of this compound. There has been no large-scale production that allows industrial processes without CO2 emissions. This is because the economic profitability of the CO2 transformation reactions strongly depends on a sustainable and renewable H2 supply in most cases, as well as on an efficient thermal and mass transport control, the implication of which is a considerable investment. Until now, such investment has not been a priority in any productive sector. However, this approach is changing because, in addition to being harmful to the environment, emitting CO2 will be somewhat more expensive thanks to the laws that regulate the limits of pollution globally, as can be seen, for instance, in the evolution of the cost of emitting CO2 in Europe (Figure 1).
Therefore, technologies are being developed to overcome the disadvantages that slow down the massive use of CO2 and one of the key points is the implementation of structured catalysts since these systems precisely favor the transport processes. In addition, with an adequate design, these allow to more effectively match the recovery of CO2 in recycling strategies of said compound in the traditional processes that emit it, increasing the global energy efficiency. These strategies agree with the new circular economy model.
In this context, it has been considered important to present a summary of the trends in policies for the control of greenhouse gas emissions and specifically CO2 (highlighting the case of Europe). How CO2 recycling is integrated into the transition to the circular economy, and the need for an important technological development that allows such integration, which is reliant on the advances on structured reactors needs to be understood.
Furthermore, the review is focused on the hydrogenation of CO2 to produce methane, known as CO2 methanation reaction or Sabatier reaction [5], because it is one of the most promising strategies within the concept of power-to-gas (P2G) [6] to convert this source of carbon in an energy vector. Although some general considerations regarding the more used catalyst for this reaction will be presented, only the application of structured systems will be outlined in detail. Surprisingly, there is a relatively small number of works devoted to the study of structured systems for CO2 methanation. These researches have begun to emerge in the last 15 years showing that this is an emerging field.

2. European Policies for the Transition to a Low-Carbon Economy as a Case of Study of the Global Panorama

Although the journey of building a comprehensive worldwide climate policy started in the 1990s, it really took off around the year 2000 with the Kyoto Protocol [7]. Nevertheless, the proposed measurements have demonstrated to be insufficient considering the critical evidence of climate change as produced by global warming [8]. This has motivated the recently formulated Paris Agreement (2015) [9] that made history by being the first globally binding agreement committed to the slowdown of global warming while seeking to reduce poverty and inequality in the world. Nevertheless, these are not easy tasks since the reduction of CO2 emissions may not be desirable in developed but principally in emerging nations. In fact, the literature presents points of view where a nexus between economic growth and CO2 emissions is observed [10,11,12]. As Mardani et al. [10] pointed out the new perspectives suggest that energy conservation policies should be appropriate to cope with the reduction of CO2 emissions without affecting economic growth.
This has motivated the formulation of the Global Goals for Sustainable Development in 2015 that the global leaders agreed to for a better world by 2030. These goals aim to fight poverty and inequality, as well as to stop climate change, thus, inspiring governments, business, civil society, and the general public to work for a better future for everyone [13]. Specifically, objective 13 (Climate Action) proposes taking urgent actions to combat climate change and its effects. This objective has launched action plans that can be an opportunity to modernize infrastructure worldwide [13]. This implies actions or targets, such as that of to integrate climate change measurements into policies and planning.
One of the clearest examples at present is that of China, which is currently positioned to become the leading economy worldwide if it continues on current projections. This implies the use of enormous resources. However, measures are being taken to prevent the effects of its great development. For instance, in 2013 it was launched the Air Pollution Action Plan [14], and probably this was China’s most influential environmental policy of the past five years [15]. Therefore, the second phase of the plan has been proposed as a new 2018–2020 Three-year Action Plan “winning the blue sky war” [15], which shows the intention of the government to improve the environmental conditions generated by development.
Unfortunately, there is not the same course of action in American policy and, taking into account, that today they are the main economy worldwide, their rejection of the Paris Agreement means the loss of one of the countries that should be leading the fight against climate change. And though the United States Environmental Protection Agency (EPA) recently showed a decrease in the United States’ Greenhouse Gas Emissions during President Trump’s first year in Office [16], it is only a matter of time before it is demonstrated that the US should resume the policies of previous governments. The European Union (EU) has confirmed its commitment with such agreement and has designed a climate policy to reduce the increase in the global average temperature to below 2 °C. For this, by 2050, the emissions of greenhouse gases (GHG) should be to 80% below the 1990 levels, and intermediate milestones for the achievement of this aim are 40% emissions cuts by 2030 and 60% by 2040 (Figure 2), according to the Low Carbon Roadmap and the Energy Roadmap [17] formulated by the EU. Both roadmaps are based primarily on economic and cost-effectiveness considerations follow criteria such as: (i) the creation of a new carbon market-based in renewable sources; (ii) the legal support for the promotion of the use of renewable energies; (iii) the technology update for enhancing the energy efficiency; (iv) the promotion of a new generation of cars and vehicles for reducing the CO2 emissions, and (v) the promotion of legal and economic support for the enhancement of technologies for the capture and store CO2 (following strategies of carbon capture and storage (CCS)), emitted by power stations and industrial buildings [18]).
However, although there is clarity in the CO2 emission targets that are expected to be reached in 2050, the transition to a low-carbon economy in basic sectors such as power generation, industry, transport, agriculture, and construction will be slightly different since each sector has its own dynamics. For instance, for the power sector, a fast transition is projected due to the probable mix of comparatively low-cost low-carbon technologies (renewable energies or sources). Similarly, the construction sector will show a rapid transition, probably because of the new design of low-energy houses, the deep renovation of existing buildings, and the increased efficiency of heating and cooling systems.
Conversely, for the transport and industry sector it is expected a moderate transition until 2030, mostly achieved through efficiency improvements, but after 2030, innovative technologies would be needed (e.g., deployment of electric mobility and Carbon Capture Storage (CCS)). Regarding the agricultural sector, a transitional low-carbon model will be closely linked to meat consumption, and hence a reduction of CO2 emissions would require behavioral changes of global population diet.
To make the transition to a low-carbon economy, it is estimated that EU should invest €270 billion, during the period 2010–2050. These investments will be in capital goods, which help the transition towards low-carbon generation technologies, such as solar, on and off-shore wind, and nuclear power, extended grid connections, new automotive and other transport technologies, low-energy houses, more efficient appliances, etc. However, despite this big economic effort, the transition to low-carbon economies will open new market opportunities, especially for the manufacturing industry once the main technological challenges are overcome. Thus, new opportunities for the achievement of a better life for the people will be created, through the generation of new jobs that will promote the economic growth of the region. In addition, the EU would become less dependent on expensive imports of oil and gas, as well as less vulnerable to the increases in oil prices. Thanks to that, the EU could save between €175–320 billion annually in fuel costs over the next 40 years [17]. Additionally, the cooperation between the member states of the EU will be strengthened through the creation of strategical alliances, such as that required for the interconnection in electricity networks. The interconnection aims to support the internal market for electricity to enable greater penetration of renewable energy and to improve the security of supply. In addition, the implementation of clean technologies and electricity-based mobility will reduce air pollution in European cities, resulting in the enhancement of the life-quality and health of the population.

3. Transition to a Circular Economy Model and Carbon Capture and Utilization (CCU) Approaches

In the last decade, the new concept and development model of the Circular Economy has been formulated as a suitable approach for facing the challenges of the current global scenario. Its aims to ensure the future of the planet by providing a better alternative to the dominant linear economic development model, the so-called “take, make and depose” [2,19] (Figure 3). In this sense, the Circular Economy model is based on three main “actions”, also known the so-called 3R’s principles: reduction, reuse, and recycle [2,20], whose aim is to address a smart use of the resources, ensuring long life cycles of products while the environmental impact of the productive processes caused by the generation of waste and pollution is reduced.
To change the economic model from a linear to a circular approach will need a modification in the way in which things are made. For example, putting sustainability and closed-loop thinking at the heart of business models and industrial organization. This kind of philosophy will have considerable implications for society [21].
From a global point of view, a circular economy could help enable developing and developed countries to industrialize, without placing unsustainable pressure on natural resources and breaching environmental limits. For companies, this kind of model offers a sustainable growth fit for a world where the prices of raw materials are higher every day.
The economic potential for this model is huge, as was projected in the McKinsey analyses carried out by the Ellen MacArthur Foundation, where it was established that if a subset of the EU manufacturing sector adopted circular economy business models, it could realize net materials cost saving worth up to €538.49 billion per year by 2025 [22]. For this economy, the resource loop would be closed, so that big volumes of finite resources, like minerals or metals, are recovered and reused [23]. Other goods can be made from plant-based materials, which are biodegradable, and can be reused as fertilizer at the end of their life. To face this, it is necessary a change in the basic structures of the industrial system.
The EU, as was commented before, has agreed to a strategy for “a resource-efficient Europe” under its strategy for 2020, and introduced an initiative to address raw-materials security [24]. Some countries have their own strategies, for example, Germany has the National Resource Efficiency Programme [25], the Netherlands proposed “materials roundabout”—a hub for the high-grade recycling of materials and products [26], and the United Kingdom has produced a study focusing on economic instruments and raw-materials security [27].
Within this context, CO2 is now one of the most abundant residues produced in linear economy models. In fact, sectors, such as chemicals, refineries, cement, iron and steel, aluminum, and paper producers, are responsible for the 68% of industry emissions [28], which represents a huge waste of raw material if CO2 becomes considered a renewable carbon source rather than a residue. This has motivated strategies such as the Carbon Capture and Utilization (CCU) that goes beyond the aims of Carbon Capture and Storage (CCS).
CCU not only presents depleting the emissions of excessive amounts of carbon to the atmosphere by capturing and storing the CO2 from the exhaust gas produced in industrial processes but also the use of such carbon dioxide as a carbon source for producing energy carriers, chemicals or materials. It is important to differentiate CCU from CCS technology, since, though both face the climate change mitigation, their impact on CO2 emissions is not the same. The potential of CCU for reversing climate change is more complex and relies on the efficiency of the entire process, the renewable character of the required energy input (considering the corresponding emissions for the production of such energy too), the storage time of CO2 in the products, and their final use [29]. In this sense, there are different sectors whose economic activity could be enhanced by the inclusion of the CO2 valorisation within their productive process, such as the petrochemical industry, the power sector, the pharmaceutical, food, polymer, stainless steel, and cement industries, among others [30,31,32,33,34,35]. Therefore, a renewed interest has been awakened in all processes that allow the chemical transformation of CO2 into value-added products.

4. CO2 Valorisation through the Sabatier Reaction

CCU technologies imply that CO2 has to be first captured [36] and then transformed into valuable products. However, although both processes deserve special attention, the present review is mainly focused on CO2 utilization rather than the capture step. In this sense, some catalytic reactions can be highlighted as possible routes for obtaining valuable products from CO2, such as olefins, methanol, formic acid, and dimethyl ether [37]. Additionally, although CO2 recovery and use may be interesting for different sectors, they must pay attention in its origin, considering parameters such as purity and the recovery procedures, for ensuring the economic profitability of the entire process. In addition, the scheme presented in Figure 4, emphasizes that H2 must be accessible to valorize CO2.
Within the presented alternatives, a preferable technology to remove and utilize the emitted CO2 gas is the catalytic conversion in methane through the Sabatier reaction, allowing a procedure to treat large amounts of CO2 in a short time [38,39]. The CO2 methanation (Equation (1)) was firstly reported by Paul Sabatier and Jean-Baptiste Senderens in 1902 [40]. Sabatier gives six main applications of the direct hydrogenation method [41], one of them is the CO hydrogenation used to produce methane or gaseous mixtures rich in methane, on a large scale. The catalytic hydrogenation of CO2 to CH4 is an old process that was mainly used for the purification of hydrogen by removing small amounts of CO2 before ammonia synthesis [42,43], and for the production of substitute natural gas (SNG) [44,45,46]. Nowadays it has received renewed interest, due to the increased global energy demand and the reduction/utilization of GHG [47,48], and particularly CO2 methanation is one of the most promising strategies within the P2G technology since the produced CH4 is a substitute for natural gas, which has an existing and well-consolidated distribution grid. Such existing distribution infrastructure would allow this resource to be exploited immediately, minimizing the investment required to massively apply this technology.
Regarding the price of natural gas, this can be cheap for some industries depending on their geographic location within in the distribution radius of a consolidated network. However, as demonstrated in Figure 1, the tendency for CO2 emissions indicates a clear growth that will force, sooner or later, industries to look for an alternative and this can be the production of methane.
However, CO2 methanation is far from being industrially implemented despite being thermodynamically favored because of the high cost of hydrogen [49,50]. Thus, the profitability of the Sabatier reaction strongly depends on the supply of cheap and renewable hydrogen.
The direct low-cost transformation of carbon dioxide into methane is highly exothermic, although less than the indirect valorization through CO methanation industrially developed [51] (Equation (2)).
CO2 methanation
CO2 + 4H2 → CH4 + 2H2O → ΔH°298K = −165 kJ · mol−1
CO methanation
CO + 3H2 → CH4 + H2O → ΔH°298K = −206 kJ · mol−1
Reverse water-gas shift (R-WGS)
CO2 + H2 ⇆ H2O + CO → ΔH°298K = 41 kJ · mol−1
The most widely accepted mechanism of this reaction is the combination of an exothermic CO methanation, and the endothermic reverse water-gas shift (R-WGS) (Equations (2) and (3) respectively) [52]. This explains the product distribution in the thermodynamic equilibrium, simulated for a stoichiometric feed-stream of CO2 (15 Vol.%) and H2 (60 Vol.%) with N2 for balance (Figure 5), where it is observable that the CO evolution is promoted with temperature resulting in the decrease of the selectivity to CH4 formation.
The main problem associated with CO2 valorization is its high chemical stability. The C=O double bond breaking is disfavored considering both entropy and enthalpy [53], and at 2000 °C only 2% dissociation into CO and O2 is estimated to occur [54].
There are recent reviews on the catalytic hydrogenation of carbon dioxide [48,55]. Supported Ru-, Rh-, or Co-based catalysts on different support oxides (TiO2, SiO2, Al2O3, CeO2, ZrO2) have been extensively studied for the hydrogenation of CO2 [56,57]. However, Ni-based catalysts remain the most widely studied materials mainly due to their low cost [58] despite the highest turnover numbers of Ru-based catalysts [59]. The process efficiency is controlled by the configuration of the catalytic reactor since its high exothermicity, but also by the catalyst activity throughout its lifetime [60]. In his comprehensive review, Jalama [59] addresses all the aspects that affect the kinetics of CO2 hydrogenation including the active phase, the support, promoters, and even pore size and particle shapes. However, this excellent study is devoted only to the hydrogenation of pure CO2.
The nature of the support plays a crucial role for high activity of CO2 methanation. This has resulted in a considerable amount of work that considers the effect of supports and promoters on the catalyst efficiency. The addition of alkaline or alkaline-earth cations promote dissociation and helps gasifying carbon deposits. The high oxygen storage capacity of reducible oxides, such as CeO2, the easiness of the reducibility of the active phase, the metal dispersion or the support basicity are key factors in the catalyst activity [56].
The CO2 methanation is sensitive to the structure of the catalysts, and this strongly depends on their chemical composition and the preparation method. Therefore, to improve the activity of the catalyst, efforts must be devoted to optimizing the synthesis conditions, the selection of high surface area supports with tailored pore sizes and optimized basicity and a careful selection of the active phase including bimetallic or multi-metallic active phases. In general, most of the studies are carried out in quite simple systems, and as a consequence, there is a strong need to develop a fundamental understanding of this reaction.
After reviewing of several works, an efficiency order of the different active phases tested for CO2 methanation can be proposed, following the sequence, Ru > Ni > Co > Fe > Mo which increases the productivity by increasing the metal content [61,62,63,64,65,66,67,68,69]. Suitable supports have been compared, result in an inferred efficiency order CeO2 > Al2O3 > TiO2 > MgO [55,68,70,71,72,73].
In the case of Ni-based catalysts, the main cause of deactivation seems to be the sintering of the metallic particles that is favored through the formation of the volatile Ni(CO)4 molecules [74]. The second cause of deactivation is the formation of carbon deposits that reduce the number of surface sites for the adsorption of CO and H2 dissociation. Resistance to sintering through a better metal-support interaction [75] and the addition of water to prevent carbon deposits [76] are common strategies, although this also influences the sintering behavior of the catalyst. Ru and other noble metals-based catalysts are stable under operating conditions and more active than Ni-based catalysts [77].
The excellent review by Jalama [59] analyzes the different aspects that influence the kinetics of the methanation reaction suggesting the general routes for CO2 hydrogenation. Two main paths are proposed for the reaction. The first one involves CO as an intermediary and a second one that does not involve CO and may have formate-like intermediates. Recent in operando spectroscopic studies mainly for Ni-based catalysts, have shown the nature of the support influences the CO2 hydrogenation mechanism [78]. Adsorbed CO and formate species are nowadays accepted to form on Ni- and Ru-based catalysts acting as active intermediate species in the hydrogenation of CO2. In some cases, CO may be formed through formate decomposition. The relative proportion of the active intermediate species is determined by the physicochemical properties of the catalyst support and, therefore, active supports do participate in the reaction mechanism.
If traces of CO2 are attempted to be converted into methane, as in the case of ammonia plants, the elevated exothermicity of the reaction does not seem to be a problem. However, in the case of the synthesis of Synthetic Natural Gas (SNG), limited heat transfer causes hot spots, affects the lifetime of construction materials and catalysts and, finally, thermal runaway of the reactor may occur. Moreover, the chemical equilibrium shifts away from optimum conditions. To avoid this, several reactor concepts have been proposed among them, cascades of fixed bed reactors with limited conversion, wall-cooled fixed bed reactors, fluidized bed reactors or slurry bubble reactors are used or currently being investigated [61]. However, very few studies are reported in process intensification for this reaction using microchannel reactors, which is an alternate concept allowing the effective reaction heat removal due to their advantageous surface-to-volume ratio [61,79]. By structuring the catalyst, the temperature hot spot is reduced to moderate values and the lifetime of the catalyst is extended by decreasing the activity loss with time.
Most of the studies are performed using pure CO2 and do not represent real industrial conditions, neither do they pay attention to the presence of contaminants. Industrial feedstocks typically contain trace amounts of sulfur compounds that in H2 atmospheres may result in H2S in the feed stream. The effect of this contaminant on the methanation of CO2 has been analyzed by Szailer et al. [80], and an unexpected result on adding small amounts of H2S (~22 ppm) to the CO2 feedstock was observed. H2S can promote the reaction on TiO2- and CeO2-supported metals clusters of Ru, Rh, and Pd. On the contrary, for ZrO2- and MgO-based supports or when the H2S content is high (~116 ppm), the catalyst results were poisoned [79]. There are, however, few studies devoted to studying the effect of these contaminants and most of them relate to CO2 extracted from biogas sources. Even fewer studies are dedicated to the analysis of flue gas methanation.
Mitsubishi disclosed a procedure for transforming the CO2 contained in power plant flue gas into methane. In their procedure, they decoupled the heat of the waste stream to be further used in the catalytic methanator loaded with Rh catalyst [81].
Müller et al. [82] tested the performances of a 60%Ni/SiO2 catalyst in the CO2 methanation reaction of conventional power plants flue gas effluents at 350 °C and stoichiometric H2-to-CO2 ratios. They report a stable catalyst with selectivity close to 100% and CO2 conversions and CH4 yields greater than 80%. However, in the presence of O2, SO2, and NO2, both CO2 conversion and CH4 yield are affected. Nitrogen oxides hardly contribute to the modification of the methanation process, and O2 up to 8% in concentration has no significant influence apart from decreasing the process efficiency due to the hydrogen combustion reaction. However, ~80 ppm SO2 concentrations result in a continuous decrease in CO2 conversion since SO2 is reduced to H2S during the reaction that poisons the nickel catalyst.
Recently, academic studies on the integration of the methanation reaction with oxycombustion system within the P2G strategy have appeared claiming the feasibility of the process [83,84,85]. Moreover, Meylan et al. [86] analyzed several scenarios for the use of CO2 emissions in the light of EU directives. They conclude that to valorize the CO2 issued from fossil resources by methanation could make sense, since it allows significant emissions savings exemplifying a possible process for the cement industries where valorization of the unavoidable CO2 emissions can help decrease industrial emissions of GHG.
Pilot plants within the P2G concept are already in operation [87]. In most cases, H2 is produced by water electrolysis, being the electrolysers powdered by renewable energy sources and almost pure CO2 after separating it from biogas using amine scrubbers. This approach with small modifications is already patented, and a scheme of the patented process is shown in Figure 6 [88].
Among them, the Audi e-gas plant, located in Werlte (Germany), has been in operation since 2013. It is based on the catalytic methanation of pure H2 and CO2 in a single isothermal fixed-bed reactor; the HELMETH project that involves Sunfire GmbH aiming to integrate high-temperature electrolysis and CO2 methanation. The Desert Research Institute (DRI) in the USA presented a system built in a trailer that consists in a PEM electrolyser powdered by solar panels and wind turbines coupled to a methanator loaded with ring-shaped PK-7R Haldor Topsoe catalyst which is able to operate up to 430 °C. Methanation of flue gases is almost exclusively attempted by research groups at the Brandenburg University of Technology who tested their approach for three months at Schwarze Pumpe power plant, in Germany [89].
A final approach that combines CO2 capture and methanation has recently been proposed. Farrauto’s group, with the support of BASF, is developing a process based on Dual Functional Materials (DFM) [90,91,92]. Their methods take the profit of the CO2 spillover from the CaO support to the metallic sites of the Ru/Al2O3 methanation catalyst [93]. However, this approach cannot be utilized using flue gases since the redox cycles imply the oxidation of metallic Ru by CO2 that is further reduced by H2. In the presence of O2, this cycle cannot take place, and the catalyst becomes inefficient [94].

5. The Use of Structured Reactors for the Sabatier Reaction

The activity and selectivity of the catalysts for these reactions are dramatically influenced by heat and mass transfer in the reactor. Heat management is a key issue in all processes involving highly exothermic or endothermic reactions, particularly when elevated conversion levels are expected, and structured catalysts may provide an optimal heat control during these processes since the high surface-to-volume ratios improve thermal management. This allows shorter space-time yields, especially in miniaturized reactors, such as those with microchannels, where the reduction of geometric dimensions allows to decrease the transport limitations. Thus, the benefits of the use of structured reactors strongly depend on their design, considering variables such as the shape, size, and density of the channels, among others related to their geometry [95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120]. Additionally, as expected, the selection of the material to manufacture the structured reactors is a determining factor [95]. Metals are highly attractive due to their high heat transfer coefficient and several cases, including alloys, have tested up to now. Most of the metallic monolithic structures developed so far are built out of Al-alloyed ferritic stainless steels foils [120,121,122,123,124,125]. However, the almost universal utilization of this alloy within the catalytic community is their use at temperatures well below the optimum service temperature and under hydrogen and carbon-rich atmospheres, resulting in migration of alloy elements to the catalytic layer, brattling its joining to the wall of the structured reactor [122,126]. Consequently, further studies about the evolution of the structured systems submitted to reactive atmospheres are required to avoid their degradation, as well as the contribution of uncontrolled processes during the reactions. What is curious is the fact that this is a subject not addressed very often and there are only a few studies in the literature that provide alternative strategies for enhancing the stability of the structured during the reactions or that present the monitoring of the degradation of the structured systems under reactive atmospheres [125,126,127,128,129]. Although this is not within the main scope of the present review, it has to be remarked that the evolution and wear of the structured during the operation have to be taken into account not only for the CO2 methanation but also for any reaction.
Even though the Sabatier reaction has been widely studied, this has been mainly carried out in fixed-bed reactors [130], and the use of structured systems has not been considered until the last decade. One of the scenarios where the advantages of the structured reactors have been attractive is the design of in-situ propellant production for Mars exploration. In particular, the Sabatier reaction is one of those considered for the design of applications able to use local resources to “live off the land”, commonly referred to as in-situ resource utilization (ISRU), which aims to expand robotic and human extraterrestrial exploration and to establish a long-term human presence beyond low earth orbit, as noted by Hu et al. [131]. These authors, in a cooperation between the US Pacific Northwest Laboratory and the NASA, proposed the conversion of Martian atmospheric CO2 into useful materials such as methane that can be used as fuel for the return journey using water electrolysis as the pathway for the production of the required hydrogen and O2 that may be used for life support. For this purpose, the use of microchannel reactors was considered since, that under this scenario, the efficiency of the process should be maximized in the so-called NASA’s ISRU. It essentially integrates a CO2 methanation unit and an R-WGS (Figure 7) but also involves complementary processes that make the process sustainable.
In this manuscript, for the Sabatier reaction, a Ru/TiO2 catalyst was used, and the structured systems were prepared using porous FeCrAlY intermetallic alloy to achieve low-pressure drop and improve heat transfer. Through single channel tests (Figure 6B), it was observed that the operation of the reactor was very steady. In general, the microreactors allowed operating under well-controlled isothermal conditions and through this work, the authors demonstrated a successful transfer of the intrinsic kinetic performance obtained at the powder catalyst level to structured systems.
More recently, within the context of the CO2 recycling strategies, Roger et al. [132] presented preliminary results of the study of structured cellular foam-based catalysts in a platelet milli-reactor. To improve the heat transfer, foams coated with different catalysts were manufactured with β-SiC. This material is chemically inert and has mechanical and thermal resistance, although it presents a low specific surface area (20 cm2/g). Nevertheless, by growing nanofibers on the foam’s surface, the specific area can be increased up to (20 cm2/g), so the effects of the nanofibers were evaluated. Later, these authors presented the performance of the open cell foams with a special focus on the exothermicity of the process, followed using infrared thermography with the experimental setup proposed presented in Figure 8 [133]. In this case, foams made with different materials were studied (alumina, aluminium, and SiC).
Through this study, it may be confirmed how important the selection of the material is. For instance, it was observable that the coating of the catalyst over every substrate requires specific treatments of the surface. Additionally, the kind of material influences the thermal profile during the reaction as well as the catalytic activity due to the differences in surface area of the substrates. The thermographs obtained during the experiments represented the temperature in the structured system and the setup allowed to establish that at the beginning of the reaction, a fast temperature increase is produced. Then the temperature increased all over the system of reaction following the flow direction. As expected, the thermal conductivity of the substrates increased in the following order: alumina < aluminium < SiC, although some hot-spots were also observed in the last case. Although it is clear that further advances are required for this experimental setup, this is an interesting approach not only for the advancement of structured catalysts but also for the integration of in-situ experiments that allow obtaining information under more realistic operational conditions.
In another work, Tada et al. [134] studied the use of sponge Ni catalysts, where the material of the structured system played the role of the active phase. The authors suggested advantages regarding the facile manufacturing of these sort of reactors. Moreover, they attributed the high activity of the devices to the great number of crystal defects of fcc-Ni in the sponge. Nevertheless, the thermal stability of the surface Ni species is the main drawback of this system since the active defects of the structured may disappear under high-temperature treatments.
Fukuhara et al. [135] reported the use of a nickel-based structured catalyst for CO2 methanation. They developed a honeycomb-type structured system with a high methanation performance. First, they tested different oxides (Al2O3, TiO2, ZrO2, Y2O3, MgO, CeO2, and zeolites) as support for the active phase (10 wt%-Ni), and the best results were obtained with Ni/CeO2, so, from this catalyst, they prepared the structured system by the wash-coating method on an aluminium-fin substrate (Figure 9). The powder catalysts exhibited a difference in the activity at 250 °C, due to the accumulation of energy as a result of the exothermicity, while the structured catalyst was able to transfer that energy excess.
Another aspect remarked by the authors is the fact that the cell density of the structured systems clearly influenced the performance. Therefore, the structured catalyst with the stacked-type-fin enhanced the performance to improve the properties of the mass transfer in the reaction field. Additionally, the stability of the structured systems was successfully probed during almost 125 h. Later in another paper derived from this work, Fukuhara and coworkers presented a more comprehensive study of the use of metallic monoliths coated with Ni/CeO2 catalyst [136]. In this case, various configurations of metallic honeycomb-type reactors, which were plain, stacked, segment, and multi-stacked (Figure 10) were constructed and evaluated under different reaction conditions (i.e., inlet temperature, feed flow rate and CO2 partial pressure), aiming to establish a suitable configuration for enhancing the heat transfer during the process.
The evaluation of the different configurations of the structured reactors revealed that multi-stacked catalysts could maintain the high methanation performance even under high feed flow rate condition of 3000 mL/min at pure feed gas component, with a moderate hot-spot formation. This resulted in the boosting of the CO2 conversion (>90%) and the CH4 selectivity (>99.5%). Furthermore, the catalytic activity was maintained at 300 °C and 3000 mL/min over 76 h. Therefore, it was confirmed that the multi-stacked configuration is a suitable approach for the enhancement of the catalytic performance during the CO2 methanation and this was associated to the enhancement of the transport phenomena according to the studies of heat and mass balances (using an overall heat transfer coefficient (U) and the reaction rate constant (K)) (Figure 11).
Probably the most comprehensive study about the heat transfer analysis during CO2 methanation using metallic structured reactors was recently presented by Schollenberger et al. [6]. In this work, stainless steel and aluminum honeycomb-type monoliths coated with Ni/Al2O3 commercial catalyst were tested and the obtained results at laboratory scale that allowed to carry out the scale-up demo-scale by means of numerical modeling. Therefore, it is important to remark that the experiments carried out in the laboratory including kinetic studies and heat transport analysis allowed to establish the basis for the technical feasibility evaluation in a bench-scale plant added to a biomass gasification plant considering the geometry and boundary conditions as presented in Figure 12.
For the heat transfer analysis, it was established the radial thermal conductivity for a honeycomb structure, strongly depends on the dimensions and the thermal conductivity of the material of construction, as well as on the gas flowing through the structure and is flow velocity. The experimental setup and the data evaluation procedure for measuring the temperature profiles in the monoliths without reaction (as reference) are presented in Figure 13, and the evaluation of the experimental results was based on the energy balance around the honeycomb represented by Equation (4). One of the most relevant results of this communication is that it was demonstrated how the thermal efficiency has to be a critical criterion for the design of a catalytic application for the CO2 methanation adaptable to a scaled-up process.
ρ g c p , g u 0 , g δ T δ z = λ r ,   eff ( δ 2 T δ r 2 + 1 r δ T δ r ) + λ z ,   eff δ 2 T δ z 2
To solve the differential equation, the following boundary conditions were set:
  • T = T i n ( r ) at z = 0 (Dirichlet boundary condition (D-BC)),
  • δ T δ z = 0 at z = L (Neumann boundary condition (N-BC)),
  • δ T δ r = 0 at r = 0 (Neumann boundary condition (N-BC)),
  • T = T w a l l ( z ) at r = R (Dirichlet boundary condition (D-BC)).
An alternative to the use of metals for manufacturing the substrates of structured reactors is the synthesis of three-dimensional network materials, controlled using synthesis. For instance, Liu et al. [137] reported the production of NiO/SBA-15 monoliths with a mesoporous framework that creates the pathways for the pass of the reactants and allows their contact with nickel species. However, heat transport is hardly achievable, despite the interesting results in this sort of structured system and considering that mass transport phenomena may be improved.
The use of ceramic monoliths in the Sabatier reaction have also been explored within the study of Ni/GDC (Ni/gadolinium-doped-ceria) structured on cordierite substrates. These devices allowed a superior heat releasing than the fixed-bed catalyst. However, the discussion was principally focused on the advantages of the use of monoliths for the enhancement of the mass transport during the reaction. Although the authors pointed out that with low-loaded monoliths (0.2 g/cm3), poor catalytic activities were observed due to the short contact time between the reactants and the catalytic layer.
Finally, the appearance of modern additive technologies for the manufacture of 3D objects, among which are the techniques of 3D printing and three-dimensional fiber deposition (3DFD) (robocasting technique), has opened a new field of exploration of design and manufacturing of multiple objects, including structured reactors. The 3DFD technique enables the production of periodic porous ceramic and metallic structures, thus, there are interesting advantages on using these techniques for obtaining structured reactors, such as the fact that more complex geometries can be produced, or that the manufacturing processes may be carried out more easily with a superior reproducibility and reducing of the production costs.
One of the few works combining these novel methodologies for the Sabatier reaction is that presented by Danaci et al. [138]. These authors proposed the manufacturing of Ni/Al2O3 coated macro-porous structures using the robocasting technique that allowed obtaining different structures made with stainless steel (316 L) (Figure 14). In general, it was confirmed the superior performance of the structured catalysts compared with that of the powder catalyst, and useful information regarding the anchoring of the catalyst to the walls of the monoliths was also presented. However, the most remarkable result is the evidence of the direct influence of the geometry design over the catalytic activity. In this case, the structured catalyst of “zig-zag” configuration showed better results and the probable improvement of the heat and mass transport phenomena must be the reason.
More recently, the same research team applied the 3DFD technique for obtaining the structured reactors using copper as raw material [139]. In all cases, the systems were coated with a Ni/Al2O3 catalyst. However, the remarkable fact of this work is that it confirms the tendency to explore different materials such as copper to optimize the obtaining of structures through 3D printing (Figure 15). This implies the optimization of different variables during the printing process as well as the during the washcoating with the selected catalyst.

6. Concluding Remarks

The current policies regarding the control of GHG emissions and the need of advances for the CO2 recycling, which is a strategy in agreement with the required transition to a circular economy model, make the application of structured reactors for CO2 methanation a topic of growing interest. In the works published up to now, it has been confirmed that the structured systems favor the mass transport during the reaction, but mainly a better transfer of heat, which is decisive for such exothermic reaction whose CH4 yields depends on a good thermal control.
Regarding the catalysts, the more active systems (mainly Ni-based catalysts), are the most widely used for manufacturing the structured catalysts. As for the manufacturing of the substrates where the catalysts are deposited, different materials have been studied. However, metals such as aluminum, copper, and stainless steels are preferred due to their high heat conduction respect to that of ceramic materials.
Despite all these positive aspects of the use of structured reactors in the Sabatier reaction, the studies presented so far do not represent a significant advance with respect to the knowledge that already exists of this type of systems in other reactions. However, the emergence of 3D printing technology has opened a wide range of possibilities in terms of the type of materials that can be used and, especially the design of the substrates, which can present geometries able to potentiate even more the benefits observed so far. These types of technologies are also allowing the development of increasingly robust simulation systems, based on reliable 3D modeling, which allows us to think that the optimization of new devices will be increasingly intelligent. Therefore, it is evident that this is a promising field in which growth in the number of works is foreseen in the near future.
Among the challenges that future works dedicated to this topic must face, is the use of materials with better conductive properties than metals. Similarly, develop designs of reactors with an optimized geometry to maximize catalytic performance and make production both cheaper and profitable. Finally, and perhaps the most difficult of all, is to study structured reactors in real applications of CO2 recycling in industrial processes that, in addition to this compound, generate another type of pollutant that can have an important impact on catalytic performance. For this, much more complex strategies are required in which the purification of CO2 before methanation will play an important role.

Funding

This research was founded by the Spanish Ministry of Economy and Competitivity through the project ENE2015-66975-C03-2-R.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ganesh, I. Conversion of carbon dioxide into methanol—A potential liquid fuel: Fundamental challenges and opportunities (a review). Renew. Sustain. Energy Rev. 2014, 31, 221–257. [Google Scholar] [CrossRef]
  2. Ghisellini, P.; Cialani, C.; Ulgiati, S. A review on circular economy: The expected transition to a balanced interplay of environmental and economic systems. J. Clean. Prod. 2016, 114, 11–32. [Google Scholar] [CrossRef]
  3. Dowell, N.M.; Fennell, P.S.; Shah, N.; Maitland, G.C. The role of CO2 capture and utilization in mitigating climate change. Nat. Clim. Chang. 2017, 7, 243. [Google Scholar] [CrossRef]
  4. EEX Group. European Emission Allowances; EEX Group: Leipzig, Germany, 2018. [Google Scholar]
  5. Stangeland, K.; Kalai, D.; Li, H.; Yu, Z. CO2 Methanation: The Effect of Catalysts and Reaction Conditions. Energy Procedia 2017, 105, 2022–2027. [Google Scholar] [CrossRef]
  6. Schollenberger, D.; Bajohr, S.; Gruber, M.; Reimert, R.; Kolb, T. Scale-Up of Innovative Honeycomb Reactors for Power-to-Gas Applications—The Project Store&Go. Chem. Ing. Tech. 2018, 90, 696–702. [Google Scholar]
  7. Kyoto Protocol—Targets for the First Commitment Period. 1997. Available online: https://ec.europa.eu/clima/policies/strategies/progress/kyoto_1_en (accessed on 31 May 2018).
  8. Klaassen, W.G.; Lefevere, J.; Meadows, D.; Slingerberg, Y.; Runge-Metzger, A.; Vergote, S.; Weksman, J.; Zapfel, P. EU Climate Policy Explained, 1st ed.; Routledge: London, UK, 2015; ISBN 978-9279482618. [Google Scholar]
  9. United Nations. “Paris Agreement” United Naitions; United Nations: New York, NY, USA, 2015. [Google Scholar]
  10. Mardani, A.; Streimikiene, D.; Cavallaro, F.; Loganathan, N.; Khoshnoudi, M. Carbon dioxide (CO2) emissions and economic growth: A systematic review of two decades of research from 1995 to 2017. Sci. Total Environ. 2019, 649, 31–49. [Google Scholar] [CrossRef] [PubMed]
  11. Zhao, X.; Zhang, X.; Li, N.; Shao, S.; Geng, Y. Decoupling economic growth from carbon dioxide emissions in China: A sectoral factor decomposition analysis. J. Clean. Prod. 2017, 142, 3500–3516. [Google Scholar] [CrossRef]
  12. Ness, D. Sustainable urban infrastructure in China: Towards a Factor 10 improvement in resource productivity through integrated infrastructure systems. Int. J. Sustain. Dev. World Ecol. 2008, 15, 288–301. [Google Scholar]
  13. United Nations. Sustainable Development GOALS. 2015. Available online: https://www.un.org/sustainabledevelopment/sustainable-development-goals (accessed on 10 June 2018).
  14. The Central People’s Government of the People’s Republic of China: Air Pollution Action Plan. 2013. Available online: http://www.gov.cn/zwgk/2013-09/12/content_2486773.htm (accessed on 10 June 2018).
  15. Eco-Business. China Releases 2020 Action Plan for Air Pollution. 2018. Available online: https://www.eco-business.com/news/china-releases-2020-action-plan-for-air-pollution (accessed on 24 June 2018).
  16. Data Shows Decrease in U.S. Greenhouse Gas Emissions during Trump’s First Year Office. 2018. Available online: http://www.epa.gov/newsreleases/data-shows-decrease-us-greenhouse-gas-emissions-during-trumps-first-year-office (accessed on 10 August 2018).
  17. 2050 Low-Carbon Economy. 2011. Available online: https://ec.europa.eu/clima/policies/strategies/2050_en#tab-0-0 (accessed on 10 August 2018).
  18. Carbon Capture and Geological Storage. 2011. Available online: https://ec.europa.eu/clima/policies/lowcarbon/ccs_en (accessed on 10 August 2018).
  19. Korhonen, J.; Honkasalo, A.; Seppälä, J. Circular Economy: The Concept and its Limitations. Ecol. Econ. 2018, 143, 37–46. [Google Scholar] [CrossRef]
  20. Figge, F.; Young, W.; Barkemeyer, R. Sufficiency or efficiency to achieve lower resource consumption and emissions? The role of the rebound effect. J. Clean. Prod. 2014, 69, 216–224. [Google Scholar] [CrossRef] [Green Version]
  21. Womack, J.P.; Jones, D.T.; Roos, D. The Machine That Changed the World; Free Press: New York, NY, USA, 1990; ISBN 13: 978-0-7432-9979-4. [Google Scholar]
  22. Fundation, E.M. Towards the Circular Economy: Accelerating the Scale-Up across Global Supply Chains; World Economic Forum: Geneva, Switzerland, 2014. [Google Scholar]
  23. Huber, J. Towards Industrial Ecology: Sustainable Development as a Concept of Ecological Modernization. J. Environ. Policy Plan. 2000, 2, 269–285. [Google Scholar] [CrossRef]
  24. Commission, E. A Resource-Efficient Europe–Flagship Initiative under the Europe 2020 Strategy; European Environment Agency: Copenhagen, Denmark, 2011. [Google Scholar]
  25. Rötggen: Germany Aims to Become World Champion in Resource Efficiency. 2011. Available online: https://www.bmu.de/en/pressrelease/roettgen-germany-aims-to-become-world-champion-in-resource-efficiency/ (accessed on 6 June 2018).
  26. Wiel, H.V.D. The Netherlands as Materials Roundabout, Waste Forum Special Edition. 2011. [Google Scholar]
  27. Hislop, H.; Hill, J.H. Reinventing the Wheel: A Circular Economy for Resource Security; Green Alliance: London, UK, 2011; ISBN 978-1-905869-46-6. [Google Scholar]
  28. Agency, I.E. Tracking Clean Energy Progress 2016; Energy Technology Perspectives: Paris, France, 2016. [Google Scholar]
  29. Onarheim, K.; Arasto, A.; Hannula, I.; Kärki, J.; Lehtonen, J.; Vainikka, P. Carbon Capture and Utilization (The Role of Carbon Capture and Utilization in Transitioning to a Low-Carbon Future; Discussion Paper; VTT: Espoo, Finland; pp. 1–21.
  30. Institute, G.C. Accelerating the Uptake of CCS: Industrial Use of Captured Carbon Dioxide. 2011. Available online: http://hub.globalccsinstitute.com/sites/default/files/publications/14026/accelerating-uptake-ccs-industrial-use-captured-carbon-dioxide.pdf (accessed on 31 August 2018).
  31. A Roadmap Por the Implementation of Carbon Utilization Technologies. CO2 Sciences and The Global CO2 Initiative Executive Summary. 2016. Available online: https://assets.ctfassets.net/xg0gv1arhdr3/27vQZEvrxaQiQEAsGyoSQu/44ee0b72ceb9231ec53ed180cb759614/CO2U_ICEF_Roadmap_FINAL_2016_12_07.pdf (accessed on 23 August 2018).
  32. US Department of Energy. Technology Program Plan. 2014. Available online: https://www.energy.gov/sites/prod/files/2017/02/f34/Ch1-SI-DOE-Energy-Technology-Roadmaps-and-Basic-Research-Needs.pdf (accessed on 23 August 2018).
  33. Styring, P.; Reith, H.D.H.; Armstrong, K. Carbon Capture and Utilization in the Green Economy. In Using CO2 to Manufacture Fuel, Chemicals and Materials; The Centre for Low Carbon Futures: New York, NY, USA, 2011; ISBN 978-0-9572588-1-5. [Google Scholar]
  34. Schüwer, D.; Bienge, A.K.K.; Viebahn, P. CO2 Reuse NRW: Evaluating Gas Sources, Demand and Utilization for CO2 and H2 within the North Rhine-Westphalia Area with Respect to Gas Qualities; Wuppertal Inst. for Climate, Environment and Energy: Wuppertal, Germany, 2015. [Google Scholar]
  35. Bocin-Dumitriu, A.; Perey Fortes, M.; Tzimas, E.; Sveen, T. Carbon Capture and Utilization Workshop, Background and Proceedings; Publications Office of the European Union: Luxembourg, 2013. [Google Scholar]
  36. Centi, G.; Perathoner, S.; Passalacqua, R.; Ampelli, C. Carbon-Neutral Fuels and Energy Carriers; CRC Press: Boca Raton, FL, USA, 2011. [Google Scholar]
  37. Online, V.A.; Centi, G.; Quadrelli, E.A.; Perathoner, S. Introduction of Renewable Energy in the value chain of chemical industries. Energy Environ. Sci. 2013, 6, 1711–1731. [Google Scholar]
  38. Pereira, M.M.; Louis, B. Carbon Dioxide, Chemical Valorization and Mitigation in the Refinery. News Future Dev. Catal. 2013, 535–562. [Google Scholar] [CrossRef]
  39. Davis, W.; Martin, M. Optimal year-round operation for methane production from CO2 and water using wind and/or solar energy. J. Clean. Prod. 2014, 80, 252–261. [Google Scholar] [CrossRef]
  40. Sabatier, P.; Senderens, J.-B. Comptes Rendus Des Seances De L’Academie Des Sciences Section VI—Chimie; Imprimerie Gauthier-Villars: Paris, France, 1902. [Google Scholar]
  41. Sabatier, P. The Method of Direct Hydrogenation by Catalysis. 1912. Available online: https://www.nobelprize.org/prizes/chemistry/1912/sabatier/lecture/A (accessed on 23 August 2018).
  42. Martin, G.A.; Primet, M.; Dalmon, J.A. Reactions of CO and CO2 on Ni/SiO2 above 373K as studied by infrarred spectroscopic and magnetic methods. J. Catal. 1978, 53, 321–330. [Google Scholar] [CrossRef]
  43. Araki, M.; Ponec, V. Methanation of carbon monoxide on nickel and nickel-copper alloys. J. Catal. 1976, 44, 439–448. [Google Scholar] [CrossRef]
  44. Garbarino, G.; Riani, P.; Busca, G. A study of the methanation of carbon dioxide on Ni/Al2O3 catalysts at atmospheric pressure. Int. J. Hydrog. Energy 2014, 39, 11557–11565. [Google Scholar] [CrossRef]
  45. Falconer, J.L.; Zagli, A.E. Adsorption and methanation of carbon dioxide on a nickel/silica catalyst. J. Catal. 1980, 62, 280–285. [Google Scholar] [CrossRef]
  46. Dalmon, J.A.; Martin, G.A. Intermediates in CO and CO2 hydrogenation over Ni catalysts. J. Chem. Soc. 1979, 75, 1011–1015. [Google Scholar] [CrossRef]
  47. Shima, M.; Sakurai, Y.; Sone, M.; Ohnishi, A.; Yoneda, T.A. Development of the Sabatier Reaction Catalyst for Practical Space Systems. Int. J. Microgravity Sci. Appl. 2013, 30, 86–93. [Google Scholar]
  48. Wang, W.; Wang, S.; Ma, X.; Gong, J. Recent advances in catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 40, 3703–3727. [Google Scholar] [CrossRef] [PubMed]
  49. Specht, M.; Baumgart, F.; Feigl, B.; Frick, V.; Stürmer, B.; Zuberbühler, U. Storing bioenergy and renewable electricity in the natural gas grid. Erdöl Erdgas Kohle 2010, 126, 342–345. [Google Scholar]
  50. Doty, F.D.; Holte, L.; Shevgoor, S. Proceedings of the ASME 2009 3rd International Conference of Energy Sustainability. In Proceedings of the ASME 2009 3rd International Conference on Energy Sustainability Collocated with the Heat Transfer and InterPACK09 Conferences, San Francisco, CA, USA, 19–23 July 2009; pp. 1–8. [Google Scholar]
  51. Bobadilla, L.F.; Muñoz-Murillo, A.; Laguna, O.H.; Centeno, M.A.; Odriozola, J.A. Does shaping catalysts modify active phase sites? A comprehensive in situ FTIR spectroscopic study on the performance of a model Ru/Al2O3 catalyst for the CO methanation. Chem. Eng. J. 2019, 357, 248–257. [Google Scholar] [CrossRef]
  52. Jürgensen, L.; Ehimen, E.A.; Born, J.; Holm-Nielsen, J.B. Dynamic biogas upgrading based on Sabatier process: Thermodynamic and dynamic process simulation. Bioresour. Technol. 2015, 178, 323–329. [Google Scholar] [CrossRef] [PubMed]
  53. Jacquemin, M.; Beuls, A.; Ruiz, P. Catalytic production of methane from CO2 and H2 at low temperature. Catal. Today 2010, 157, 462. [Google Scholar] [CrossRef]
  54. Weingberg, W.H. Why CO2 Does not Dissociate on Rh at Low Temperature. Surf. Sci. 1983, 128, 231–235. [Google Scholar]
  55. Aziz, M.A.A.; Jalil, A.A.; Triwahyono, S.; Ahmad, A. CO2 methanation over heterogeneous catalysts: Recent progress and future prospects. Green Chem. 2015, 17, 2647–2663. [Google Scholar] [CrossRef]
  56. Tejada, L.M.M.; Muñoz, A.; Centeno, M.A.; Odriozola, J.A. In situ Raman spectroscopy study of Ru/TiO2 catalyst in the selective methanation of CO. J. Raman Spectrosc. 2016, 47, 189–197. [Google Scholar] [CrossRef]
  57. Razzaq, R.; Li, C.S.; Usman, M.; Suzuki, K.; Zhang, S.J. A highly active and stable Co4N/gamma-Al2O3 catalyst for CO and CO2 methanation to produce synthetic natural gas (SNG). Chem. Eng. J. 2015, 262, 1090–1098. [Google Scholar] [CrossRef]
  58. Xu, X.; Moulijn, J.A. Mitigation of CO2 by Chemical Conversion: Plausible Chemical Reactions and Promising Products. Energy Fuel 1996, 10, 305–325. [Google Scholar] [CrossRef]
  59. Jalama, K. Carbon dioxide hydrogenation over nickel-, ruthenium-, and copper-based catalysts: Review of kinetics and mechanism. Catal. Rev. 2017, 59, 95–164. [Google Scholar] [CrossRef]
  60. Kopyscinski, J.; Schildauer, T.J.; Biollaz, S.M.A. Production of synthetic natural gas (SNG) from coal and dry bomass—A technology review from 1950 to 2009. Fuel 2010, 89, 1765–1783. [Google Scholar] [CrossRef]
  61. Türks, D.; Mena, H.; Armsbruster, U.; Martin, A. Methanation of CO2 on Ni/Al2O3 in a Structured Fixed Bed Reactor—A Scale-Up Study. Catalysts 2017, 7, 152. [Google Scholar] [CrossRef]
  62. Du, G.; Lim, S.; Yang, Y.; Wang, C.; Pfefferle, L.; Haller, G.L. Methanation of carbon dioxide on Ni-incorporated MCM-41 catalysts: The influence of catalyst pretreatment and study of steady-state reaction. J. Catal. 2007, 249, 370–379. [Google Scholar] [CrossRef]
  63. Kuśmierz, M. Kinetic study on carbon dioxide hydrogenation over Ru/γ-Al2O3 catalysts. Catal. Today 2008, 137, 429–432. [Google Scholar] [CrossRef]
  64. Ma, S.; Tan, Y.; Han, Y. Methanation of syngas over coral reef-like Ni/Al2O3 catalysts. J. Nat. Gas Chem. 2011, 20, 435–440. [Google Scholar] [CrossRef]
  65. Narayan, R.L.; King, T.S. Hydrogen adsorption states on silica-supported Ru–Ag and Ru–Cu bimetallic catalysts investigated via microcalorimetry. Thermochim. Acta 1998, 312, 105–114. [Google Scholar] [CrossRef]
  66. Pastor-Pérez, L.; Baibars, F.; le Sache, E.; Arellano-García, H.; Gu, S.; Reina, T.R. CO2 valorisation via Reverse Water-Gas Shift reaction using advanced Cs doped Fe-Cu/Al2O3 catalysts. J. CO2 Util. 2017, 21, 423–428. [Google Scholar] [CrossRef]
  67. Schanke, D.; Vada, S.; Blekkan, E.A.; Hilmen, A.M.; Hoff, A.; Holmen, A. Study of Pt-Promoted Cobalt CO Hydrogenation Catalysts. J. Catal. 1995, 156, 85–95. [Google Scholar] [CrossRef]
  68. Lapidus, A.L.; Gaidai, N.A.; Nekrasov, N.V.; Tishkova, L.A.; Agafonov, Y.A.; Myshenkova, T.N. The mechanism of carbon dioxide hydrogenation on copper and nickel catalysts. Petrol. Chem. 2007, 47, 75–82. [Google Scholar] [CrossRef]
  69. Janke, C.; Duyar, M.S.; Hoskins, M.; Farrauto, R. Catalytic and adsorption studies for the hydrogenation of CO2 to methane. Appl. Catal. B 2014, 153, 184–191. [Google Scholar] [CrossRef]
  70. Baronskiy, M.; Rastorguev, A.; Zhuzhgov, A.; Kostyukov, A.; Krivoruchko, O.; Snytnikov, V. Photoluminescence and Raman spectroscopy studies of low-temperature γ-Al2O3 phases synthesized from different precursors. Opt. Mater. 2016, 53, 87–93. [Google Scholar] [CrossRef]
  71. Garbarino, G.; Bellotti, D.; Riani, P.; Magistri, L.; Busca, G. Methanation of carbon dioxide on Ru/Al2O3 and Ni/Al2O3 catalysts at atmospheric pressure: Catalysts activation, behaviour and stability. Int. J. Hydrogen Energy 2015, 40, 9171–9182. [Google Scholar] [CrossRef]
  72. Gupta, N.M.; Kamble, V.S.; Iyer, R.M.; Thampi, K.R.; Gratzel, M. The transient species formed over Ru-RuOx/TiO2 catalyst in the CO and CO + H2 interaction: FTIR spectroscopic study. J. Catal. 1992, 137, 473–486. [Google Scholar] [CrossRef]
  73. Centi, G.; Perathoner, S. Opportunities and prospects in the chemical recycling of carbon dioxide to fuels. Catal. Today 2009, 148, 191–205. [Google Scholar] [CrossRef]
  74. Agnelli, M.; Swaan, H.M.; Marquez-Alvarez, C.; Martin, G.A.; Mirodatos, C. CO hydrogenation on a nickel-catalyst -2- a mechanistic study by transient kinetics and infrared-spectroscopy. J. Catal. 1998, 175, 117–128. [Google Scholar] [CrossRef]
  75. Nguyen, T.T.M.; Wissing, L.; Skjoth-Rasmussen, M.S. High temperature methanation: Catalyst considerations. Catal. Today 2013, 215, 233–238. [Google Scholar] [CrossRef]
  76. Gao, J.; Wang, Y.; Ping, Y.; Hu, D.; Xu, G.; Gu, F.; Su, F. A thermodynamic analysis of methanation reactions of carbon oxides for the production of synthetic natural gas. RSC Adv. 2012, 2, 2358–2368. [Google Scholar] [CrossRef]
  77. Kustov, L.; Frey, A.M.; Larsen, K.E.; Johannessen, T.; Norskov, J.K.; Christensen, C.H. CO methanation over supported bimetallic Ni-Fe catalysts: From computational studies towards catalyst optimization. Appl. Catal. A 2007, 320, 98–104. [Google Scholar] [CrossRef]
  78. Aldana, P.A.U.; Ocampo, F.; Kobl, K.; Louis, B.; Thibault-Starzyk, F.; Daturi, M.; Bazin, P.; Thomas, S.; Roger, A.C. Catalytic CO2 Valorization into CH4 on Ni-based ceria-zirconia. Reaction mechanism by operando IR spectroscopy. Catal. Today 2013, 215, 201–207. [Google Scholar] [CrossRef]
  79. Brooks, K.P.; Hu, J.; Zhu, H.; Kee, R.J. Methanation of carbon dioxide by hydrogen reduction using the Sabatier process in microchannel reactors. Chem. Eng. Sci. 2007, 62, 1161–1170. [Google Scholar] [CrossRef]
  80. Szailer, T.; Novak, E.; Oszko, A.; Erdohelyi, A. Effect of H2S on the hydrogenation of carbon dioxide over supported Rh catalysts. Top. Catal. 2007, 46, 79–86. [Google Scholar] [CrossRef]
  81. Bergins, C.; Buddenberg, T.; Kakaras, E. Methanation Method and Power Plant c Omprising CO2 Methanation of Power Plant Flue Gas. EP2014/064,625, 15 January 2015. [Google Scholar]
  82. Müller, K.; Fleige, M.; Rachow, F.; Schmeiber, D. Sabatier CO2-methanation of flue gas emitted by conventional power plants. Energy Procedia 2013, 40, 240–248. [Google Scholar] [CrossRef]
  83. Meylan, F.D.; Moreau, V.; Erkman, S. CO2 utilization in the perspective of industrial ecology, an overview. J. CO2 Util. 2015, 12, 101–108. [Google Scholar] [CrossRef]
  84. Bailera, M.; Lisbona, P.; Romeo, L.M.; Espatolero, S. Power to Gas-biomass oxycombustion hybrid system: Energy integration and potential applications. Appl. Energy 2016, 167, 221–229. [Google Scholar] [CrossRef]
  85. Gutierreaz-Martin, F.; Rodriguez-Anton, L.M. Power-to-SNG technology for energy storage at large scales. Int. J. Hydrogen Energy 2016, 41, 19290–19303. [Google Scholar] [CrossRef]
  86. Meylan, F.D.; Piguet, F.P.; Erkman, S. Power-to-gas through CO2 methanation: Assessment of the carbon balance regarding EU directives. J. Energy Storage 2017, 11, 16–24. [Google Scholar] [CrossRef]
  87. Bailera, M.; Lisbona, P.; Romeo, L.M.; Espatolero, S. Power to Gas projects review: Lab, pilot and demo plants for storing renewable energy and CO2. Renew. Sustain. Energy Rev. 2017, 69, 292–312. [Google Scholar] [CrossRef]
  88. Schildhauer, T.J.; Elber, U.; Nachtegaal, M.; Gubler, L.; Janshon, P. Integrated Process/Plant for Storage of CO2 by Conversion to Synthetic Natural Gas; P.S. Institute: San Francisco, CA, USA, 2015. [Google Scholar]
  89. Integrated High-Temperature Electrolysis and Methanation for Effective Power to Gas Conversion; Project ID 621210; Helmet—Report Summary; Karlsruher Institut Fuer Technologie: Karlsruhe, Germany, 2017.
  90. Duyar, M.S. A Study of Catalytic Carbon Dioxide Methanation Leading to the Development of Dual Function Materials for Carbon Capture and Utilization; University of Columbia: New York, NY, USA, 2015. [Google Scholar]
  91. Duyar, M.S.; Farrauto, R.J.; Park, A.-A. Methods Systems and Materials for Capturing Carbon Dioxide and Converting it to a Chemical Product; University of Columbia: New York, NY, USA, 2016. [Google Scholar]
  92. Zheng, Q.; Farrauto, R.; Nguyen, A.C. Adsorption and Methanation of Flue Gas CO2 with Dual Functional Catalytic Materials: A Parametric Study. Ind. Eng. Chem. Res. 2016, 55, 6768–6776. [Google Scholar] [CrossRef]
  93. Duyar, M.S.; Treviño, M.A.A.; Farrauto, R.J. Dual function materials for CO2 capture and conversion using renewable H2. Appl. Catal. B Environ. 2015, 168, 370–376. [Google Scholar] [CrossRef]
  94. Wang, S.; Schrunk, E.T.; Mahajan, H.; Farrauto, R.J. The Role of Ruthenium in CO2 Capture and Catalytic Conversion to Fuel by Dual Function Materials (DFM). Catalysts 2017, 7, 88. [Google Scholar] [CrossRef]
  95. Cybulski, A.; Moulijn, J.A. Chapter 1: The present and the future of structured catalysts: An overview. In Structured Catalysts and Reactors; Cybulski, A., Moulijn, J.A., Eds.; Taylor and Francis Group: Boca Raton, FL, USA, 2006; pp. 1–16. [Google Scholar]
  96. Groppi, G.; Tronconi, E. Design of novel monolith catalyst supports for gas/solid reactions with heat exchange. Chem. Eng. Sci. 2000, 55, 2161–2171. [Google Scholar] [CrossRef]
  97. Groppi, G.; Tronconi, E. Simulation of structured catalytic reactors with enhanced thermal conductivity for selective oxidation reactions. Catal. Today 2001, 69, 63–73. [Google Scholar] [CrossRef]
  98. Reymond, J.P. Structured supports for noble catalytic metals: Stainless steel fabrics and foils, and carbon fabrics. Catal. Today 2001, 69, 343–349. [Google Scholar] [CrossRef]
  99. Wunsch, R.; Fichtner, M.; Gorke, O.; Haas-Santo, K.; Schubert, K. Process of applying Al2O3 coatings in microchannels of completely manufactured microstructured reactors. Chem. Eng. Technol. 2002, 25, 700–703. [Google Scholar] [CrossRef]
  100. Goerke, O.; Pfeifer, P.; Schubert, K. Water gas shift reaction and selective oxidation of CO in microreactors. Appl. Catal. A Gen. 2004, 263, 11–18. [Google Scholar] [CrossRef]
  101. Jähnisch, K.; Hessel, V.; Lowe, H.; Baerns, M. Chemistry in microstructured reactors. Angew. Chem.-Int. Ed. 2004, 43, 406–446. [Google Scholar] [CrossRef] [PubMed]
  102. Kolb, G.; Hessel, V. Micro-structured reactors for gas phase reactions. Chem. Eng. J. 2004, 98, 1–38. [Google Scholar] [CrossRef]
  103. Cao, C.S.; Wang, Y.; Rozmiarek, R.T. Heterogeneous reactor model for steam reforming of methane in a microchannel reactor with microstructured catalysts. Catal. Today 2005, 110, 92–97. [Google Scholar] [CrossRef]
  104. Kiwi-Minsker, L.; Renken, A. Microstructured reactors for catalytic reactions. Catal. Today 2005, 110, 2–14. [Google Scholar] [CrossRef] [Green Version]
  105. Groppi, G.; Baretta, A.; Tronconi, E. Monolithic Catalysts for Gas-Phase Syntheses of Chemicals. In Structured Catalysts and Reactors; Cybulski, A., Jacobs, G., Eds.; CRC Press—Taylor & Francis Group: Boca Ratón, FL, USA, 2006; pp. 243–310. [Google Scholar]
  106. Twigg, M.T.; Webster, D.E. Chapter 3: Metal and coated metal catalyts. In Structured Catalysts and Reactors; Cybulski, A., Moulijn, J.A., Eds.; Taylor and Francis Group: Boca Raton, FL, USA, 2006; pp. 71–105. [Google Scholar]
  107. Graça, I.; Gonzáles, L.V.; Bacariza, M.C.; Frenandes, A.; Henriques, C.; Lopes, J.M.; Riberio, M.F. CO2 hydrogenation into CH4 on NiHNaUSY zeolites. Appl. Catal. B Environ. 2014, 147, 101–110. [Google Scholar] [CrossRef]
  108. Snajdrová, V.; Hlincík, T.; Ciahotny, K.; Polak, L. Pilot unit of carbon dioxide methanation using nickel-based catalysts. Chem. Pap. 2018, 9, 2339–2346. [Google Scholar]
  109. Kolb, G.; Baier, T.; Schuerer, J.; Tiemann, D.; Ziogas, A.; Specchia, S.; Galletti, C.; Germani, G.; Schuurman, Y. A micro-structured 5 kW complete fuel processor for iso-octane as hydrogen supply system for mobile auxiliary power units—Part II—Development of water-gas shift and preferential oxidation catalysts reactors and assembly of the fuel processor. Chem. Eng. J. 2008, 138, 474–489. [Google Scholar] [CrossRef]
  110. Pennemann, H.; Hessel, V.; Kolb, G.; Loewe, H.; Zapf, R. Partial oxidation of propane using micro structured reactors. Chem. Eng. J. 2008, 135, S66–S73. [Google Scholar] [CrossRef]
  111. Kolb, G.; Hofmann, C.; O’Connell, M.; Schuerer, J. Microstructured reactors for diesel steam reforming, water-gas shift and preferential oxidation in the kiloWatt power range. Catal. Today 2009, 147, S176–S184. [Google Scholar] [CrossRef]
  112. Kolb, G.; Schelhaas, K.P.; Wichert, M.; Burfeind, J.; Hesske, C.; Bandlamudi, G. Development of a Micro-Structured Methanol Fuel Processor Coupled to a High-Temperature Proton Exchange Membrane Fuel Cell. Chem. Eng. Technol. 2009, 32, 1739–1747. [Google Scholar] [CrossRef]
  113. Renken, A.; Kiwi-Minsker, L. Microstructured Catalytic Reactors. In Advances in Catalysis; Gates, B.C., Knozinger, H., Jentoft, F.C., Eds.; Academic Press: Cambridge, MA, USA, 2010; pp. 47–122. [Google Scholar]
  114. Kolb, G.; Keller, S.; Pecov, S.; Pennemann, H.; Zapf, R. Development of Micro-structured Catalytic Wall Reactors for Hydrogen Production by Methanol Steam Reforming over Novel Pt/In(2)O(3)/Al(2)O(3) Catalysts. In Proceedings of the 10th International Conference on Chemical and Process Engineering, Florence, Italy, 8–11 May 2011; pp. 133–138. [Google Scholar]
  115. Almeida, L.C.; Sanz, O.; Merino, D.; Arzamendi, G.; Gandía, L.M.; Montes, M. Kinetic analysis and microstructured reactors modeling for the Fischer–Tropsch synthesis over a Co–Re/Al2O3 catalyst. Catal. Today 2013, 215, 103–111. [Google Scholar] [CrossRef]
  116. Kolb, G. Review: Microstructured reactors for distributed and renewable production of fuels and electrical energy. Chem. Eng. Process. 2013, 65, 1–44. [Google Scholar] [CrossRef]
  117. Reyero, I.; Velasco, I.; Sanz, O.; Montes, M.; Arzamendi, G.; Gandia, L.M. Structured catalysts based on Mg-Al hydrotalcite for the synthesis of biodiesel. Catal. Today 2013, 216, 211–219. [Google Scholar] [CrossRef]
  118. Palma, V.; Miccio, M.; Ricca, A.; Meloni, E.; Ciambelli, P. Monolithic catalysts for methane steam reforming intensification: Experimental and numerical investigations. Fuel 2014, 138, 80–90. [Google Scholar] [CrossRef]
  119. Qiao, J.; Liu, Y.; Hong, F.; Zhang, J. A review of catalysts for the electroreduction of carbon dioxide to produce low-carbon fuels. Chem. Soc. Rev. 2014, 43, 631–675. [Google Scholar] [CrossRef] [PubMed]
  120. Almeida, L.C.; Echave, F.J.; Sanz, O.; Centeno, M.A.; Odriozola, J.A.; Montes, M. Washcoating of metallic monoliths and microchannel reactors. In Studies in Surface Science and Catalysis; Gaigneaux, E.M., Devillers, M., Hermans, S., Jacobs, J.A., Eds.; Elsevier: Amsterdam, The Netherlands, 2010; pp. 25–33. [Google Scholar]
  121. Sanz, O.; Echave, F.; Romero-Sarria, F.; Odriozola, J.A.; Montes, M. Advances in Structured and Microstructured Catalytic Reactors for Hydrogen Production, Renewable Hydrogen Technologies; Gandia, L.M., Arzamendi, G., Dieguez, P.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2013; pp. 201–224. [Google Scholar]
  122. Laguna, O.H.; Domínguez, M.I.; Centeno, M.A.; Odriozola, J.A. Chapter 4: Catalysts on metallic surfaces: Monoliths and microreactors. In New Materials for Catalytic Applications; Parvulescu, V.I., Kemnitz, E., Eds.; Elsevier: Amsterdam, The Netherlands, 2016; pp. 81–120. [Google Scholar]
  123. Tejada, L.M.M.; Sanz, O.; Dominguez, M.I.; Centeno, M.A.; Odriozola, J.A. AISI 304 Austenitic stainless steels monoliths for catalytic applications. Chem. Eng. J. 2009, 148, 191–200. [Google Scholar]
  124. Tejada, L.M.M.; Dominguez, M.I.; Sanz, O.; Centeno, M.A.; Odriozola, J.A. Au/CeO2 metallic monolith catalysts: Influence of the metallic substrate. Gold Bull. 2013, 46, 221–231. [Google Scholar] [CrossRef]
  125. Domínguez, M.I.; Perez, A.; Centeno, M.A.; Odriozola, J.A. Metallic structured catalysts: Influence of the substrate on the catalytic activity. Appl. Catal. A Gen. 2014, 478, 45–57. [Google Scholar] [CrossRef]
  126. Laguna, O.H.; Domínguez, M.I.; Centeno, M.A.; Odriozola, J.A. Forced deactivation and postmortem characterization of a metallic microchannel reactor employed for the preferential oxidation of CO (PROX). Chem. Eng. J. 2016, 302, 650–662. [Google Scholar] [CrossRef]
  127. Heid, B.; Frischat, G.H.; Helimold, P. Silicate enamels for stainless steel. In Proceedings of the XXI International Enamellers Congress, Shangai, China, 18–22 May 2008; pp. 68–74. [Google Scholar]
  128. Serres, T.; Dreibine, L.; Schuurman, Y. Synthesis of enamel-protected catalysts for microchannel reactors: Application to methane oxidative coupling. Chem. Eng. J. 2012, 213, 31–40. [Google Scholar] [CrossRef]
  129. Shieu, F.S.; Deng, M.J.; Lin, K.C.; Wong, J.C.; Wu, J.Y. Effect of surface pretreatments on the adherence of porcelain enamel to a type 316L stainless steel. J. Mater. Sci. 1999, 34, 5265–5272. [Google Scholar] [CrossRef] [Green Version]
  130. Younas, M.; Kong, L.L.; Bashir, M.J.K.; Nadeem, H.; Shehzad, A.; Sethupathi, S. Recent Advancements, Fundamental Challenges, and Opportunities in Catalytic Methanation of CO2. Energy Fuel 2016, 30, 8815–8831. [Google Scholar] [CrossRef]
  131. Hu, J.; Brooks, K.P.; Holladay, J.D.; Howe, D.T.; Simon, T.M. Catalyst development for microchannel reactors for martian in situ propellant production. Catal. Today 2007, 125, 103–110. [Google Scholar] [CrossRef]
  132. Frey, M.; Édouard, D.; Roger, A.-C. Optimization of structured cellular foam-based catalysts for low-temperature carbon dioxide methanation in a platelet milli-reactor. C. R. Chim. 2015, 18, 283–292. [Google Scholar] [CrossRef]
  133. Frey, M.; Romero, T.; Roger, A.-C.; Edouard, D. Open cell foam catalysts for CO2 methanation: Presentation of coating procedures and in situ exothermicity reaction study by infrared thermography. Catal. Today 2016, 273, 83–90. [Google Scholar] [CrossRef]
  134. Tada, S.; Ikeda, S.; Shimoda, N.; Honma, T.; Takahashi, M.; Nariyuki, A.; Satokawa, S. Sponge Ni catalyst with high activity in CO2 methanation. Int. J. Hydrogen Energy 2017, 42, 30126–30134. [Google Scholar] [CrossRef]
  135. Fukuhara, C.; Hayakawa, K.; Suzuki, Y.; Kawasaki, W.; Watanabe, R. A novel nickel-based structured catalyst for CO2 methanation: A honeycomb-type Ni/CeO2 catalyst to transform greenhouse gas into useful resources. Appl. Catal. A Gen. 2017, 532, 12–18. [Google Scholar] [CrossRef]
  136. Ratchahat, S.; Sudoh, M.; Suzuki, Y.; Kawasaki, W.; Watanabe, R.; Fukuhara, C. Development of a powerful CO2 methanation process using a structured Ni/CeO2 catalyst. J. CO2 Util. 2018, 24, 210–219. [Google Scholar] [CrossRef]
  137. Liu, Q.; Tian, Y. One-pot synthesis of NiO/SBA-15 monolith catalyst with a three-dimensional framework for CO2 methanation. Int. J. Hydrogen Energy 2017, 42, 12295–12300. [Google Scholar] [CrossRef]
  138. Danaci, S.; Protasova, L.; Lefevere, J.; Bedel, L.; Guilet, R.; Marty, P. Efficient CO2 methanation over Ni/Al2O3 coated structured catalysts. Catal. Today 2016, 273, 234–243. [Google Scholar] [CrossRef]
  139. Danaci, S.; Protasova, L.; Snijkers, F.; Bouwen, W.; Bengaouer, A.; Marty, P. Innovative 3D-manufacture of structured copper supports post-coated with catalytic material for CO2 methanation. Chem. Eng. Process. 2018, 127, 168–177. [Google Scholar] [CrossRef]
Figure 1. Evolution of the allowance for the emission of CO2 in Europe (Adapted from the European Emissions Allowances data reported in Reference [4]).
Figure 1. Evolution of the allowance for the emission of CO2 in Europe (Adapted from the European Emissions Allowances data reported in Reference [4]).
Catalysts 08 00578 g001
Figure 2. Transition to a low-carbon European Union (EU) economy in 2050 (Green House Gas (GHG) emissions by sector over time as % of 1990 levels). Adapted from [17].
Figure 2. Transition to a low-carbon European Union (EU) economy in 2050 (Green House Gas (GHG) emissions by sector over time as % of 1990 levels). Adapted from [17].
Catalysts 08 00578 g002
Figure 3. Difference between linear economy and circular economy.
Figure 3. Difference between linear economy and circular economy.
Catalysts 08 00578 g003
Figure 4. Possible catalytic reactions for the CO2 valorization.
Figure 4. Possible catalytic reactions for the CO2 valorization.
Catalysts 08 00578 g004
Figure 5. Thermodynamic equilibrium analysis of CO2 methanation (H2/CO2 = 4/1 molar; P = 1 atm).
Figure 5. Thermodynamic equilibrium analysis of CO2 methanation (H2/CO2 = 4/1 molar; P = 1 atm).
Catalysts 08 00578 g005
Figure 6. An overview of the synergetic effects of combining diverse industrial processes. Adapted from Reference [88].
Figure 6. An overview of the synergetic effects of combining diverse industrial processes. Adapted from Reference [88].
Catalysts 08 00578 g006
Figure 7. The NASA’s in-situ resource utilization (ISRU): (A,B) Top view of the reactor channel installed with structured catalyst; (C) Catalyst stability testing under repeated shutting down and restart conditions (3% Ru/TiO2 as structured catalysts). Adapted from Hu et al. [131].
Figure 7. The NASA’s in-situ resource utilization (ISRU): (A,B) Top view of the reactor channel installed with structured catalyst; (C) Catalyst stability testing under repeated shutting down and restart conditions (3% Ru/TiO2 as structured catalysts). Adapted from Hu et al. [131].
Catalysts 08 00578 g007
Figure 8. Experimental setup of the study of the performance of the platelet milli-reactor Adapted from Frey et al. [133].
Figure 8. Experimental setup of the study of the performance of the platelet milli-reactor Adapted from Frey et al. [133].
Catalysts 08 00578 g008
Figure 9. Image of the honeycomb-type Ni-based catalysts and catalytic performance of the honeycomb-type Ni/CeO2 catalyst with different cell densities. Adapted from Fukuhara et al. [135].
Figure 9. Image of the honeycomb-type Ni-based catalysts and catalytic performance of the honeycomb-type Ni/CeO2 catalyst with different cell densities. Adapted from Fukuhara et al. [135].
Catalysts 08 00578 g009
Figure 10. Configuration of the honeycomb-type catalysts. Reproduced with permission from Reference [136]. Copyright S. Ratchahat, M. Sudoh, Y. Suzuki, W. Kawasaki, R. Watanabe, C. Fukuhara, 2018.
Figure 10. Configuration of the honeycomb-type catalysts. Reproduced with permission from Reference [136]. Copyright S. Ratchahat, M. Sudoh, Y. Suzuki, W. Kawasaki, R. Watanabe, C. Fukuhara, 2018.
Catalysts 08 00578 g010
Figure 11. (A) Enhancement of the heat and mass transport phenomena with the stacked and segment-type configurations: (A) U values; (B) K values. Reproduced with permission from Reference [136]. Copyright S. Ratchahat, M. Sudoh, Y. Suzuki, W. Kawasaki, R. Watanabe, C. Fukuhara, 2018.
Figure 11. (A) Enhancement of the heat and mass transport phenomena with the stacked and segment-type configurations: (A) U values; (B) K values. Reproduced with permission from Reference [136]. Copyright S. Ratchahat, M. Sudoh, Y. Suzuki, W. Kawasaki, R. Watanabe, C. Fukuhara, 2018.
Catalysts 08 00578 g011
Figure 12. Reactor geometry and boundary conditions for numerical modeling. Reproduced with permission from Reference [6]. Copyright D. Schollenberger, S. Bajohr, M. Gruber, R. Reimert, T. Kolb, 2018.
Figure 12. Reactor geometry and boundary conditions for numerical modeling. Reproduced with permission from Reference [6]. Copyright D. Schollenberger, S. Bajohr, M. Gruber, R. Reimert, T. Kolb, 2018.
Catalysts 08 00578 g012
Figure 13. Experimental setup for the study of honeycomb-type monoliths: position of the monolith in the measuring section (left), the positioning of the thermocouples (middle), and a schematic volume element for balancing (right). Reproduced with permission from Reference [6]. Copyright D. Schollenberger, S. Bajohr, M. Gruber, R. Reimert, T. Kolb, 2018.
Figure 13. Experimental setup for the study of honeycomb-type monoliths: position of the monolith in the measuring section (left), the positioning of the thermocouples (middle), and a schematic volume element for balancing (right). Reproduced with permission from Reference [6]. Copyright D. Schollenberger, S. Bajohr, M. Gruber, R. Reimert, T. Kolb, 2018.
Catalysts 08 00578 g013
Figure 14. Monoliths obtained using robocasting for the Sabatier reaction. (A) Cross-sectional images of the different configuration of the structured monoliths obtained using three-dimensional fiber deposition (3DFD); (B) Catalytic performance of the structured reactors vs. the powder catalyst. Adapted from Danaci et al. [138].
Figure 14. Monoliths obtained using robocasting for the Sabatier reaction. (A) Cross-sectional images of the different configuration of the structured monoliths obtained using three-dimensional fiber deposition (3DFD); (B) Catalytic performance of the structured reactors vs. the powder catalyst. Adapted from Danaci et al. [138].
Catalysts 08 00578 g014
Figure 15. 3D printed structured reactors made with copper. Reproduced with permission from Reference [139]. Copyright S. Danaci, L. Protasova, F. Snijkers, W. Bouwen, A. Bengaouer, P. Marty, 2018.
Figure 15. 3D printed structured reactors made with copper. Reproduced with permission from Reference [139]. Copyright S. Danaci, L. Protasova, F. Snijkers, W. Bouwen, A. Bengaouer, P. Marty, 2018.
Catalysts 08 00578 g015

Share and Cite

MDPI and ACS Style

Navarro, J.C.; Centeno, M.A.; Laguna, O.H.; Odriozola, J.A. Policies and Motivations for the CO2 Valorization through the Sabatier Reaction Using Structured Catalysts. A Review of the Most Recent Advances. Catalysts 2018, 8, 578. https://doi.org/10.3390/catal8120578

AMA Style

Navarro JC, Centeno MA, Laguna OH, Odriozola JA. Policies and Motivations for the CO2 Valorization through the Sabatier Reaction Using Structured Catalysts. A Review of the Most Recent Advances. Catalysts. 2018; 8(12):578. https://doi.org/10.3390/catal8120578

Chicago/Turabian Style

Navarro, Juan C., Miguel A. Centeno, Oscar H. Laguna, and José A. Odriozola. 2018. "Policies and Motivations for the CO2 Valorization through the Sabatier Reaction Using Structured Catalysts. A Review of the Most Recent Advances" Catalysts 8, no. 12: 578. https://doi.org/10.3390/catal8120578

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop