Next Article in Journal
Applications of Radial Basis Functional Link Networks in the Exploration for Lala Copper Deposits in Sichuan Province, China
Previous Article in Journal
Editorial for Special Issue “Water within Minerals Processing”
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Prospects of Clay Minerals from the Baltic States for Industrial-Scale Carbon Capture: A Review

by
Jānis Krūmiņš
,
Māris Kļaviņš
*,
Rūta Ozola-Davidāne
and
Linda Ansone-Bērtiņa
Department of Environmental Science, Faculty of Geography and Earth Sciences, University of Latvia, Raiņa Blvd. 19, LV-1586 Riga, Latvia
*
Author to whom correspondence should be addressed.
Minerals 2022, 12(3), 349; https://doi.org/10.3390/min12030349
Submission received: 18 February 2022 / Revised: 11 March 2022 / Accepted: 13 March 2022 / Published: 14 March 2022
(This article belongs to the Section Clays and Engineered Mineral Materials)

Abstract

:
Carbon capture is among the most sustainable strategies to limit carbon dioxide emissions, which account for a large share of human impact on climate change and ecosystem destruction. This growing threat calls for novel solutions to reduce emissions on an industrial level. Carbon capture by amorphous solids is among the most reasonable options as it requires less energy when compared to other techniques and has comparatively lower development and maintenance costs. In this respect, the method of carbon dioxide adsorption by solids can be used in the long-term and on an industrial scale. Furthermore, certain sorbents are reusable, which makes their use for carbon capture economically justified and acquisition of natural resources full and sustainable. Clay minerals, which are a universally available and versatile material, are amidst such sorbents. These materials are capable of interlayer and surface adsorption of carbon dioxide. In addition, their modification allows to improve carbon dioxide adsorption capabilities even more. The aim of the review is to discuss the prospective of the most widely available clay minerals in the Baltic States for large-scale carbon dioxide emission reduction and to suggest suitable approaches for clay modification to improve carbon dioxide adsorption capacity.

1. Introduction

Carbon dioxide (CO2) is one of the major greenhouse gasses (GHGs), which, if released into the atmosphere, is reflecting infrared radiation back to the Earth’s surface and traps the radiant heat. Therefore, excessive CO2 emissions are of global concern as they are amongst the main contributors to climate change. Furthermore, global trends of CO2 emissions show annual growth, which is then naturally followed by an annual increase in the average temperature. Recent data indicate that in October 2021, the increase of CO2 concentration was 2.46 ± 0.26 ppm y−1, but according to the National Oceanic and Atmospheric Administration (NOAA) 2020 annual climate report, the temperature (land and ocean combined) has increased at an average rate of 0.08 °C per decade since 1980 [1,2]. The consequences thereof are gradual warming and drying of the climate that, among other threats, is causing major and devastating wildfires worldwide, which in fact release massive amounts of CO2 to the atmosphere themselves and make carbon emissions into even a larger issue [3]. In reality, an ever-increasing CO2 concentration in the atmosphere has a number of other secondary effects on the environment, such as changes in the hydrogeological cycle, increased occurrence of various extreme climate events, sea-level rise, species migration, harvest loses, increased occurrence of infectious diseases, and others [4,5,6,7,8].
The evidence suggests that the atmospheric CO2 concentration has increased from the pre-industrialization level (1750) of 280 ppm to 413.3 ppm in 2021 [9,10]. Furthermore, the rate of CO2 growth currently is rapidly speeding up and, according to the future predictions, CO2 concentration will continue to increase and can reach 670 ppm by year 2100 if no action is taken as soon as possible [9]. Essentially, far-reaching measures are required to deal with this growing climate change emergency. It is crucial that forward-looking measures to reduce the impact of climate change caused by increasing CO2 concentration include alternative, green, and sustainable energy sources. In addition, incentives to increase energy efficiency are required as indicated in recent political declarations of the European Union (EU), United Nations (UN), and other international organisations [11,12,13].
Given that the replacement of energy sources is a slow, gradual, and yet expensive process, measures to absorb industrially produced CO2 emissions are amidst current primary tasks. At the same time, actions for energy source transition must also be implemented immediately, in parallel with emission reduction. The most sustainable strategies to limit these CO2 emissions is carbon capture and storage (CCS) and, where it is possible, the subsequent CO2 utilization [14].
The climate change mitigation plan includes carbon capture to address the emissions mostly coming from burning fossil fuel in the energy production and transport sectors, which is then followed by carbon storage, for example, in the geological environment or industrially designed media. In addition, where possible, the utilization of stored CO2 as an ingredient or component for practical applications is also supported and advisable. In fact, the development of new approaches on how to use CO2 as a source material for other substances, therefore transforming this waste product into a valuable resource for other products, is also among modern topical issues that requires immediate solution [15,16,17,18].
Depleted oil and gas reservoirs, or hydrocarbon reservoirs in general, are viewed as one of the top options for CCS technologies; this is because these structures are sealed in by thick layers of hard impervious rocks, which can prevent possible CO2 leakage. In fact, long-term structural integrity of these sealing structures is a precondition and imperative for avoiding leakage from pools of hydrocarbons [19,20,21]. Therefore, alongside research of these reservoirs for potential CO2 storage, much attention has also been paid to their sealing geological structures.
The impervious sedimentary rocks, which are isolating hydrocarbon reservoirs, often contain high percentages of various clay minerals, such as kaolinite, smectite, illite, and others, thus their potential in relation to CCS has also been evaluated worldwide. At the same time, it should be noted that finding suitable storage reservoirs or media is only half of the solution for the climate crisis. The other half of the solution includes sustainable applications and technology for carbon capture directly from the emission sources that could be used on an industrial scale. Therefore, this review article addresses possibilities for carbon capture by clay minerals and their prospective for use in applications and technologies for CO2 emission reduction, including approaches for clay modification to improve the efficiency of use.
Clay minerals are hydrous aluminium (Al3+) phyllosilicates. Besides Al3+, they often contain variable amounts of Fe2+, Fe3+, alkali metals (Na+, K+), alkaline earths (Mg2+, Ca2+), and other cations. Clay minerals are common weathering products and one of the end results of the low-temperature hydrothermal alteration in rocks. Therefore, they are common in soil and in fine-grained sedimentary and metamorphic rocks [22]. In regard to hydrocarbon reservoir sealing structures, the most common constituents at depths up to 2–3 km and temperature up to 100 °C are smectite clay minerals, among which the most common smectite mineral is montmorillonite [23]. Beyond this depth range and temperature limit, smectites go through the diagenesis and transform into illite clay minerals [24]. Illite comparatively has lower cation-exchange capacity (CEC) than smectite and, in addition, it is also non-expanding clay, whereas smectite is expanding clay and thus can be considered as much more valuable for a variety of uses, including CO2 adsorption. Moreover, given that most hydrocarbon reservoirs with any reasonable potential to store CO2, due to pressure and temperature limitations, are within the depth of 2–3 km, smectite interaction with CO2 has been of special scientific attention for decades [24,25,26,27,28]. In addition, smectite and other clay minerals have been viewed in regard to CO2 capture not only as the part of sealing structures around hydrocarbon reservoirs, but also individually as materials with reasonable CO2 adsorption capacity, which can be used in applications and technologies to trap CO2 from the atmosphere on an industrial level [20,24,28]. Unfortunately, there are major drawbacks of CO2 sequestration attempts, which reside in the low overall concentration in the main CO2 emission sources. To overcome these drawbacks, scientists worldwide seek new and alternative technologies that among all also encompass a variety of clay mineral modification approaches [29]. In this context, the review summarizes the prospective of clay minerals for carbon capture on an industrial scale and suggests strategies for clay modification to improve CO2 adsorption capacity.
Although this review focuses on clay mineral assemblages that are available in the Baltic States to explore the potential of sustainable and cost-effective local natural resource usage, the authors believe that the information provided is comprehensive and representative of clay applications on a global scale to reach the carbon capture aims.

2. Global and Regional Carbon Dioxide Emissions

Global CO2 emissions are the primary driver of climate change. It is universally recognized that to avoid irreversible damage to the climate, the civilization immediately needs to reduce GHG emissions. If no action is taken as soon as possible, climate change will fundamentally reshape the society and geopolitical relations; in perspective, it will destroy economies and health care systems. This is a global and comprehensive issue, which everyone must be involved in tackling. Unfortunately, due to a lack of public and political support for specific pro-environmental policies, it has been an endless point of contention in international discussions with strong resistance against immediate action [30,31,32].
The majority of anthropogenic CO2 emissions are produced in urban areas, which therefore is a challenge for a sustainable development. It was expected that during the COVID-19 pandemic in 2020 and 2021, with continuation in 2022, when due to worldwide lockdowns the industry was significantly less active, CO2 emissions would decrease [33]. However, the decrease did not stand out from any natural CO2 variations observed before and the average growth rate kept increasing (Table 1) [34]. In addition, global wildfires that seem to increase annually by the extent of area they occupy were producing perhaps similar if not higher CO2 emissions than any possible decrease resulting from this pandemic.
Most of the anthropogenic CO2 emissions on the planet are produced in the Northern Hemisphere, with the largest emitters being Eastern Asia, Western Europe, and the Northeast of North America, where GHG is a major by-product of industrial activities. In general, five global economic sectors with the highest produced CO2 emissions are:
  • Transportation (predominantly non-commercial motor vehicles);
  • Electricity production (fossil fuel burning);
  • Industry (manufacturing, processing, etc.);
  • Residential (heating, cooking, etc.);
  • Agriculture (farm equipment, agricultural machinery, etc.).
Although these sectors have varied in correlative sequence depending on the given state economy, emissions within these sectors fundamentally reside in fossil fuel burning. Even though the agriculture is commonly seen as a non-fuel contributor of CO2 emissions, the agricultural machinery, such as harvesters, tractors, and other technologies are clearly a significant contributor of fossil fuel CO2 emissions, especially in top agricultural producing countries. In view of this, the largest producers of fossil fuel CO2 emissions worldwide in 2019 were China, United States (US), India, and Russia [35]. Even though statistics for 2020 at the time of this review had not yet been compiled, predictable trends were similar to those before.
Trends of the annual atmospheric CO2 emissions typically are measured at the Mauna Loa observatory. Mauna Loa is an island of Hawaii, isolated in the middle of the Pacific Ocean at over 3 km above sea level. The CO2 measurements are performed at the upper north face of Mauna Loa volcano, where there is no impact from any industrial objects or forests that may cause increase or decrease in CO2 concentration within their vicinity. The CO2 sensors are positioned so they can sample an incoming breeze directly from the ocean, unaffected by any factors on the island [36]. Recent measures indicated that the monthly average Mauna Loa atmospheric CO2 in September 2021 was 413.30 ppm, which was nearly 2 ppm increase from the year before, when in September 2020 it was 411.52 ppm [10]. Even the worldwide suspension of industry in 2020 due to COVID-19 did not yield any significant reduction in CO2 emissions and trends keep showing annual increase, which therefore requires an immediate action to slow down this trend. In 2020, global CO2 emissions declined by 5.8%, which is the largest decline ever recorded; however, global energy-related CO2 emissions remained high and contributed to CO2 emissions reaching their highest ever average concentration in the atmosphere at 412.5 ppm at the end of the year, followed by yet another increase in 2021 [37]. The only option to slow down climate change is to halt the annual increase of CO2 emissions, which requires the implementation of carbon neutrality policies, prudent use of fossil resources, and transition to environmentally friendly energy sources. The future of the Earth requires creating and sustaining an effective strategic dialogue among governments, industries, and societies worldwide.
Even though the carbon emissions vary greatly from country to country, reducing these emissions must be a global effort. As for the CO2 emission level in the Baltic States, it is determined by the range of energy sources of the region. The range, however, differs significantly among the three Baltic States (Table 2) and thus the emissions there are also contrasting. Nevertheless, all three countries are gradually making greater use of renewable resources, especially in electricity production, where such energy sources as solar, wind, hydropower, biomass, and others are put into practice, and Latvia in this respect is the leader with 61.11% share of low-carbon sources in the electricity production.
Furthermore, the Baltic States have ratified the Paris Climate Agreement and have defined their strategic climate targets for emission reduction; this among all includes majorly reducing the use of fossil resources (oil, coal, and gas) in energy production (Table 3) [38,39].
While the share of energy sources differs amongst the Baltic States (Table 3), all three countries are highly dependent on energy imports. In this regard, Estonia has the lowest share of imports due to its access to domestic oil shale resources [39]. In comparison, in Latvia a major part of the total primary energy supply comes from firewood, but in Lithuania until 2009, more than 70% of primary supply was provided by nuclear energy (zero emission energy). The last Lithuanian nuclear power plant (Ignalina), however, was decommissioned in 2009, due to its similarities to the Chernobyl power plant [39]. Differences in the use of energy sources are reflected in distinct CO2 emissions amongst Baltic countries. For instance, according to Eurostat (last update on 17 August 2021), in 2019 the Estonian energy sector produced 11,975.96 tCO2, Latvia 6975.52 tCO2, but Lithuania caused the release of 11,240.45 tCO2 into the atmosphere [40].

3. The Technologies for Carbon Dioxide Capture and Storage

The ever-increasing atmospheric CO2 concentration calls for immediate action to reduce possibilities for global catastrophic risks to the planet and to the very existence of civilization. This is why the development of technologies for CCS are crucial for preserving the future and the quality of life. During the last decades, the development of CCS technologies has been studied widely and intensively to find the balance among efficiency of use, production, and maintenance costs. Notably, the advancement of technologies for carbon capture is equally as important as the new options for carbon storage. The reduction of CO2 emissions generally relies on its separation from flue gas from the combustion reactions. The technologies for CO2 separation include six major methods, which are [41]:
  • Absorption;
  • Adsorption;
  • Calcium looping;
  • Cryogenic separation;
  • Membrane separation;
  • Biological separation with microalgae.
The CO2 absorption technologies generally can be distinguished between chemical and physical absorption. The chemical absorption is then subdivided into formulation and operation. The chemical formulation involves a CO2 reaction with single amines (derivatives of ammonia), amine blends (solutions of two or more amines in varying compositions), and caustics (strong alkaline chemicals). At the same time, chemical operation is based on the use of a chemical adsorber or stripper and rotating columns to absorb CO2, where the rotation increases the chemical reactivity and the absorption efficiency [42]. The second absorption type, physical absorption of CO2, involves the use of units with various solvents and sorbents for gas separation, such as selexol, rectisol, fluor (econamine), purisol, and others [41,43,44,45,46].
Technologies for CO2 adsorption can be distinguished between adsorbent beds and the use of regeneration cycles. The fixed-bed adsorption generally utilizes activated carbon, alumina (synthetically produced Al2O3), silica, zeolites, metal organic frameworks (MOFs), hydrotalcites, amine supported adsorbents, and polymers [41,47,48,49,50,51,52,53,54]. Regeneration cycles, in turn, involve pressure swing adsorption, temperature swing adsorption, steam, or moisture adsorption [41,55,56,57].
The calcium looping technology uses the regenerative calcium cycle, where to separate CO2 from flue gas, a metal is reversibly reacted between its carbonate form and its oxide form. Within this approach for CO2 separation, the technology basically involves a loop of calcination and carbonation to remove CO2 from flue gas [41,58].
The cryogenic technology includes a set of conventional and unconventional methods. Among these, the cryogenic distillation is amidst the most common conventional approaches. In turn, most unconventional methods involve various cryogenic fluids, heat exchangers, and cryogenic packed beds to fix CO2 [41,59,60,61].
The membrane technology is based on gas separation using specially designed membranes, for instance, a polyphenylene oxide (PPE), poly dimethyl siloxane (PDMS) membrane, or polypropylene (PP) or ceramic bed systems to fix CO2 [41,62,63,64].
The technology behind the biological CO2 separation using microalgae involves carbon bio-fixation, for example, through enzyme-catalytic hydrolysis. The idea behind microbial bio-fixation involves CO2 fixation for microalgae cell growth [41,65,66,67].
Different technologies require divergent approaches; they are also characterized by different use and maintenance costs, as well as the range of applications and scale. Currently, among main technologies used for CO2 capture is [68,69,70,71,72,73]:
  • Cryogenic distillation;
  • Membrane purification;
  • Electrochemical separation;
  • Absorption with liquids;
  • Adsorption using solid materials.
Unfortunately, most of these approaches are expensive and high-energy demanding, which therefore limit widespread use of the given technology.
The cryogenic distillation is amongst most used technologies for CO2 adsorption and gas separation [69]. It is low-temperature CO2 capture technology, which relies on phase change. The technology can be applied to a range of CO2 concentrations. Regrettably, it is high-energy demanding and expensive, and thus the use on large industrial scale cannot be economically justifiable. At the same time, this technology can be recommended for smaller-scale applications [41].
The membrane purification involves CO2 capture in polymeric membranes. The membrane purification technology requires comparatively less energy than cryogenic distillation and is also much easier to operate; however, this technique is limited to minor CO2 concentrations and thus again cannot be justified for an industrial scale use [70,74].
The electrochemical reduction of CO2 is among promising future technologies; however, there are numerous challenges for large-scale production [75]. Separation of CO2 from flue gas using electrochemical cells is complex and energy demanding process. In this context, experiments have shown that this technology can be used to separate CO2 from flue gas streams produced by coal combustion for electricity generation; however, for this to happen, elevated temperature molten carbonate electrochemical cells would be required and the presence of any kind of contaminants would impact the electrolyte within the cell and thus also the effectiveness of CO2 separation. At the same time, experiments with low-temperature cells also continue and their applications are continuously advancing, which show comparatively higher work efficiency than the high-temperature cells [76].
Yet another approach—CO2 adsorption with liquids, in general with alkylamine solution or chilled ammonia, collides with an obstacle—is a complex, expensive, and high-energy demanding technique. Moreover, evaporation of these liquids eventually causes equipment corrosion, and thus this method may not be among long-term solutions [77].
In many aspects, the CO2 adsorption using solids is among the most reasonable options as this method demands the least energy, can be used in the long-term, and some of the materials can be reused and applied on a large industrial scale. In addition, organic or inorganic sorbents are available with high selectivity in respect to CO2, high efficiency, and stability during their exploitation [73,78,79]. Use of porous solid materials, such as clays, zeolites, activated carbon, MOFs, metal oxides, layered double hydroxides (LDHs), and porous silica have shown to be very effective for selective CO2 capture [80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95]. Notably, the majority of CO2 adsorption studies have been reported in high pressure and elevated temperatures as it is known that with increasing pressure, CO2 adsorption level would also increase [96]. However, in regard to GHG removal from the atmosphere, CO2 adsorption at atmospheric pressure also has to be accounted for, and therefore it is crucial to expand on this topic in further studies [97].
The most common technologies for CO2 capture from large stationary sources are generally based on a selective CO2 absorption using aqueous solutions of amines, such as monoethanolamine (MEA), diethanolamine (DEA), N-methyldiethanolamine (MDEA), diglycol-amine (DGA), and 2-amino-2-methyl-1-propanol (AMP) [98]. Regrettably, such an approach has several disadvantages over other equivalent solutions for carbon capture; for example, some known disadvantages are equipment corrosion or loss of amines due to evaporation. Moreover, solvent regeneration requires high energy consumption, and thus alternative approaches that are less energy demanding are necessary. In fact, this necessity has promoted the development of new strategies for CO2 capture that also involve the development of new appropriate solid material sorbents based on cheap substances, for instance, on clay minerals.

4. Materials for Carbon Dioxide Capture Technologies

It is reasonable to consider that several groups of sorbents, both representing solid and liquid phase materials, are suggested for the development of the CCS technologies due to their high efficiency and selectivity to CO2 capture, long-term applicability, and economic justifiability. In this regard, the materials with high potential for CO2 capture include a variety of porous materials and liquids of both natural and artificial origin, such as [99,100,101,102,103]:
  • Metal organic frameworks (MOFs);
  • Graphene organic frameworks (GOFs);
  • Covalent organic frameworks (COFs);
  • Metal oxides;
  • Homogenous porous silica;
  • Zeolites;
  • Activated carbon;
  • Clay minerals;
  • Molecular basket sorbents (MBSs);
  • Ionic liquids.
It has been reported that MOFs and GOFs form three-dimensional structures with narrow and homogenous pore size distribution, and this allows reversible retention of CO2 molecules [99,100,101,104,105]. A number of MOFs with designed pore and channel size and large surface area are viewed as materials for CO2 sorption and storage [72,106].While MOFs and GOFs show significant advantages in gas selectivity and separation in comparison to traditional adsorbents, the main disadvantages of these structures, however, are their relatively low thermal and hydrolytic stability, which are crucial for adsorbent regeneration and low yields. Yet another obstacle is the excessive cost, which is required to prepare these materials on a large industrial scale [99]. At the same time, properties of MOFs can be further improved by several means [107,108]. The main improvements that can be made to MOF composites are increased porosity and sorption capacity, as well as improved special functionality. Moreover, numerous structural modifications and improved kinetics in the synthesis of MOFs can be made, that all together can significantly improve their effectiveness towards CO2 sorption [109].
Other structures—COFs—are an equal alternative to MOFs and GOFs, as they can be either two- or three-dimensional highly porous organic solids, and they have already been used in the photocatalytic and electrocatalytic systems for large scale CO2 conversion to CO. However, such photocatalytic systems have poor overall CO2 selectivity and low effectiveness for CO2 adsorption [110,111,112,113,114,115]. Moreover, COF applications on an industrial scale are hampered by their low stability in the presence of water [116,117]. Simultaneously, studies show an effective CO2 adsorption on numerous metal oxides, such as Fe2O3, TiO2, ZrO2, or Al2O3 in the temperature below 0 °C, but unfortunately, the high energy support to maintain the required temperature is high, making this approach high-energy demanding [118]. Notably, the use of such solids as CaO, MgO, or mixed oxides coming from hydrotalcites generally is of low cost; they are highly available and have high overall capacity to capture CO2. However, use of these oxides is affected by the same complications as other metal oxides [119].
In the last several decades, experimentation with porous silica in regard to CO2 adsorption has been carried out. Experiments with porous silica with different textural parameters such as mesocellular foams—MCM-41, MCM-48, SBA-15, hexagonal mesoporous silica, and others—have shown that CO2 capacity is directly proportional to the microporosity [119,120,121]. However, in comparison, MOFs show much better CO2 adsorption capacity than silica [90,99,120,122,123,124,125].
Zeolites and activated carbon, among all materials, reach by far the highest CO2 adsorption values. In fact, inorganic zeolites were the first porous materials to be investigated for the ability to sequester CO2; however, their production is expensive, thus alternatives with similar properties are being sought [99,126,127,128,129]. Zeolites, activated carbon, and other solid porous sorbents are capable to capture CO2 via physical and chemical adsorption mechanisms [98]. Zeolites can bind CO2 either by chemical binding or by including both chemical and physical sorption pathways. In comparison, clay minerals are a low-cost alternative with similar CO2 adsorption capacity to zeolites and activated carbon, thus their use on an industrial scale can be economically justifiable. Clays have high pore and channel size variability, large surface area, a wide variety of structures responsible for CO2 sorption, and large void volume for carbon storage. Furthermore, clays can be chemically modified to ensure high clay-based sorbent selectivity in respect to other gasses present in flue gas, such as H2N2, O2, CH4, or CO [72,99]. For instance, Wang et al. (2013) have reported that the support material accounts for over 70% of the total capital cost for sorbent preparation. In this respect, due to the easy availability, excellent thermal and chemical stability, surface properties that can be adjusted for various properties, and low cost, clay minerals can be considered one of the most justifiable materials for CCS [130]. For instance, one of the top sequesters for CO2 is the exchanged fluorohectorite clay, which belongs to the smectite group [131]. In addition, clay minerals can be used as the base material for other technologies, such as MBSs. The concept of MBS is used to denote sorbents that may selectively capture CO2 onto a functional basket. Sorbents of this type are typically prepared by immobilizing an amine-functional polymer onto a porous carrier. Recent studies point to the benefits of MBS, such as, superior sorption–desorption characteristics, high sorption capacity, selectivity towards CO2, high regeneration abilities, and high stability. In addition, MBSs have much less corrosion potential and require much lower energy consumption when compared to the conventional methods, such as amine scrubbing [132]. It is worth mentioning that, while sorption processes have been studied in high detail and general regularities are known, desorption processes are not yet fully understood. At the same time, recent studies show that CO2 desorption in natural clays and clay modification products, including MBSs, can be controlled both by pore structures and by chemically functionalized groups [133]. However, understanding which factor prevails requires further investigation. Nevertheless, it is believed that desorption can be triggered by changes in clay properties that may endure pressure or temperature change, yet further studies are required.
The last considerable group of materials with a potential for CO2 fixation are ionic liquids. These are non-volatile liquids with modifiable structure and high CO2 uptake capacity [134]. Ionic liquids have been explored in various chemical and biological applications; however, use of them comes with a number of potential weaknesses in comparison to clay minerals; this includes weak thermodynamic properties, corrosivity, and toxicity [135]. One of the least substantial options for industrial-scale carbon capture is the precipitation of calcium carbonate in underground reservoirs. Crystallization of calcium carbonate requires a liquid, such as groundwater, thus this method can be viewed within the ionic fluids. However, the crystallization is slow geological process that is sensitive to pressure and temperature changes, and thus this approach is not an effective long-term or industrial-scale solution [136]. Therefore, the use of clay minerals and clay modification to support an increased level of CO2 fixation are both justifiable and welcomed in regard to reducing GHG emissions.

5. Structural and Property Features of Clay Minerals to Support Carbon Dioxide Capture

The ability of clay minerals to fix CO2 relies on the structural and property features of a particular clay mineral. While two broad categories of clays can be distinguished (cationic and anionic clays), only one type of these clays can be widely found in nature. Cationic clays are clay minerals found in nature, but anionic clays are LDHs. While cationic clays are widespread in nature, anionic clays are often synthesised due to their rarity and thus their extraction from natural sources cannot be justifiable on a major scale [137]. In addition, the cationic clays often have negatively charged alumina-silicate (Al2O3-SiO4) layers, with small cations in the interlayer space to balance the charge. In comparison, the anionic clays have positively charged brucite (Mg(OH)2) type metal hydroxide layers with balancing ions and water molecules, which are located interstitially [137,138]. Therefore, there are noticeable structural differences between both clay types and therefore also in their support for CO2 capture.
Cationic clay minerals are hydrous Al3+ or Mg2+ phyllosilicates with two-dimensional layered structure [139]. In addition, according to the Association Internationale pour l’Etude des Argiles (AIPEA) and the Clay Minerals Society (CMS) nomenclature committee, clay minerals impart plasticity when wet and harden upon drying or firing [140]. This means that clay minerals generally are water-containing compounds with structure in which silicate tetrahedrons are arranged in sheets. Moreover, clay mineral properties change depending on the water content within. Each layer of clay mineral structure consists of a silica tetrahedral sheet and a metal oxide or hydroxide octahedral sheet, which are linked together in certain proportions [141,142,143]. The above-mentioned silica tetrahedral sheet consists of the central silicon (Si4+) cation, which is surrounded by four O2 atoms (Figure 1). This sheet is formed when three out of four O2 atoms, known as basal oxygens, are corner-linked with three nearest tetrahedrons; this in turn forms an infinite two-dimensional hexagonal mesh pattern [141,144]. The metal oxide or hydroxide octahedral sheet consists of a central cation, which most often is a Al3+ or Mg2+ cation that is surrounded by six O2 or hydroxyl (OH) atoms (Figure 1). This sheet is formed when neighbouring octahedrons edge-share two O2 atoms with each other and the smallest structural unit contains three such octahedrons and is of a hexagonal or pseudohexagonal symmetry [141,142,144,145].
Cationic clay minerals form due to weathering and decomposition of igneous rocks that are in contact with air, water, or steam, and thus they are very common in soils, sedimentary rocks, metamorphic rocks, and volcanoclastic rocks [22,146,147]. While there are numerous clay minerals, they are generally classified into nine groups, which are based on the variations in the chemical composition and atomic structure of these minerals. These groups are [148,149]:
  • Kaolin-serpentine (kaolinite, nacrite, dickite etc.);
  • Pyrophyllite-talc;
  • Mica;
  • Vermiculite;
  • Smectite (montmorillonite);
  • Chlorite;
  • Sepiolite-palygorskite;
  • Interstratified clay minerals;
  • Allophane-imogolite.
The key factor for determining clay mineral type essentially is the SiO2 ratio in a formula [146]. In addition, clay minerals are also often classified by layer types or tetrahedral and octahedral sheet combinations, which allows to distinguish three general clay mineral types [150]:
  • 1:1 layer type, tetrahedral–octahedral sheet combination;
  • 2:1 layer type, tetrahedral–octahedral–tetrahedral sheet combination;
  • 2:1:1 layer type, tetrahedral–octahedral–tetrahedral sheet combination.
Yet another clay mineral classification considers the charge per formula unit, dioctahedral, or trioctahedral character and chemical composition of the current clay mineral type. Based on both these criteria, there are 10 main clay mineral groups [151,152,153]:
  • Serpentine (e.g., chrysolite, lizardite);
  • Kaolin (e.g., kaolinite, dickite, halloysite);
  • Talc;
  • Pyrophyllite;
  • Smectite (e.g., saponite, hectorite, montmorillonite, beidellite);
  • Vermiculite;
  • True mica (e.g., illite, glauconite, phlogopite, biotite);
  • Brittle mica (e.g., clintonite, margarite);
  • Chlorite (e.g., clinochlore, chamosite, donbassite);
  • Mixed layer group (e.g., chlorite-smectite, chlorite-vermiculite, illite-smectite).
Essentially, clay minerals are constructs of two-dimensional tetrahedral stacks of inorganic layers, where atom substitutions generate a negative charge on each layer surface of clay mineral, which is then balanced by exchangeable interlayer cations [28,154]. These interlayer cations determine clay physicochemical characteristics [155,156,157]. The structural and composition features of particular clay minerals give them different physical and chemical properties, e.g., particle size and shape, specific surface area, CEC, plasticity, hydration, swelling, and surface electric charge [28,154,155,156,157,158]. The overall structural arrangement of clay minerals is simple; however, the complexity arises due to the wide variety of isomorphic substitutions that can occur within the aluminosilicate layers, the disordered nature of the interlayer region, stacking defects in the clay layer sequence, or the variable interlayer separation [154].
In regard to the CO2 capture, minerals such as kaolin, smectite, and brittle mica group have been investigated as prospective materials with necessary properties for carbon capture [99].
The structure of kaolin group minerals (1:1 layer type) is composed of one alumina octahedral sheet and one silica tetrahedral sheet, which are stacked one above other (Figure 2a,b) [141,159]. These sheets are kept together by H2 bridges between O2 atoms on the tetrahedral sheet and surface hydroxyl groups on the octahedral sheet [141,142]. The sheet has little or no permanent charge due to the low amount of substitution. As a result, kaolinite has rather low CEC and small surface area [141,142]. In addition, kaolinite is a triclinic (three-dimensional geometrical arrangement) mineral with pseudohexagonal crystals that arrange in book-like aggregates, and their particle size ranges from 0.05 μm to 2 μm [142,144].
The structure of smectite clay minerals (2:1 layer type) is composed of two silica tetrahedral sheets that encompass one octahedral sheet, which contains Al3+, Fe3+, or Mg2+ cations (Figure 2d,e) [143,159,160]. One of the characteristic features of smectite minerals is their isomorphous substitution in the octahedral and tetrahedral sheets. There is a substitution of trivalent for tetravalent ions (Al3+ for Si4+) in the tetrahedral sheet and of divalent for trivalent ions (Mg2+ for Al3+) in the octahedral sheet [152,158,160]. The substitution results in a charge deficiency, which is then balanced by hydrated cations (K+, Na+, Ca2+ and Mg2+) in the interlayer site, which is exchangeable, and thus this leads to high CEC [144,152,159]. The particle size of smectite minerals usually does not exceed 0.5 μm, while the shape varies from subhedral lamellae with irregular outlines to euhedral lamellae with rhombic outlines [144]. The most common smectite minerals are Na+ and Ca2+ montmorillonites, whose layer charge deficiency is balanced by the interlayer cations, such as sodium or calcium, and water molecules [158].
Fibrous clay minerals, such as palygorskite (attapulgite) and sepiolite, have a unique arrangement of the tetrahedral and octahedral sheets that are distinct from the typical 1:1 and 2:1-layer types. Palygorskite and sepiolite are composed of continuous two-dimensional silica tetrahedral sheets and one Mg2+ octahedral sheet, which is continuous in only one dimension. As a result, the tetrahedral sheets are split into ribbons and each of these ribbons are linked to the next ribbon by inversion of silica tetrahedral (Figure 2c). The dimension of channels between the ribbon strips is ranging from 4 to 6 angstroms for palygorskite and from 4 to 9.5 angstroms for sepiolite [141,143,158]. Both minerals are of fibrous morphology with micropores of around 0.54 nm in diameter; at the same time, mineral crystals are arranged in the bundles in an elongated shape. Notably, these properties are highly varied in clay minerals and even more so within different fibrous clay minerals. The morphology of these clay minerals is the most important physical attribute, which determines a high surface area and provides more active sites and thus also high potential for CO2 sorption [158,161].
The interaction between CO2 and clay minerals has attracted the interest of the scientific community as such minerals have proven potential to capture carbon [20,24,28]. In addition, clay minerals are a class of potential adsorbents that also possess good stability and economic viability [162]. Moreover, clay minerals are universally available and versatile materials, and due to their physicochemical properties, they can also be used as ion exchangers or catalysts [138]. Therefore, the use cationic or anionic clay minerals, in addition to liquid and supported amines, are amongst options viewed for trapping CO2 [81,163,164,165,166,167,168]. In this regard, cationic montmorillonite due to its slight basicity seems to be one of the best options for carbon capture [169,170,171,172,173,174,175,176]. In addition, numerous studies suggest that clay minerals can also be used as support for immobilizing enzymes. These enzymes can be immobilized onto clay minerals through non-covalent adsorption and covalent bonding [139]. The non-covalent adsorption involves van der Walls forces, electrostatic interactions, hydrogen bonding, and hydrophobic interactions. The covalent bonding, in turn, can be used to avoid possible desorption of enzymes [139].
In terms of clay mineral parameters with high importance on CO2 sorption capacity, the moisture may be among the most important ones. This is because there is almost always water in CO2 containing gas mixtures. Moreover, it is worth noting that the sorption mechanism of CO2, for example, on amine containing sorbents differs significantly in humid and in dry conditions. For example, in humid conditions between one CO2 and two amine groups, bicarbonates are formed, but in dry conditions carbamates are forming instead. Thus, sorbed amount of CO2 in humid conditions is comparatively higher than in dry conditions [177]. This was proven by Irani et al. (2015), who reported that the addition of 1 vol% of water could increase CO2 sorption capacity from 2.2 mmol g−1 in dry conditions to 3.8 mmol g−1 in humid conditions. Yuan et al. (2018) had similar findings, revealing that the adsorption capacity increased from 1.93 mmol g−1 in dry conditions to 2.21 mmol g−1 in humid conditions. At the same time, there were indications that adsorption–desorption cycles show good regenerability in both dry and humid conditions [177,178].

6. Clay Mineral and Carbon Dioxide Interaction Depending on Mineral Type

Studies of CO2 sorption onto clay minerals started more than 50 years ago with aims to analyse the surface area of clay minerals and to identify the sorption mechanism [179,180]. Nowadays, clay minerals and CO2 interactions are studied with two main aims: one is the use of clay minerals as sorbents for CO2 capture, and another is the geological storage of captured CO2 by clay minerals or in the underground structures containing clay minerals [96,99,181,182]. Additionally, common clay mineral types have been used to study their potential for gas separation and carbon sequestration processes. In fact, by numerous analytical methods, it has also been suggested to characterise CO2 sorption process, stability of formed complexes, and desorption processes [24,99,183,184,185]. For example, smectite and CO2 interaction has been investigated by X-ray diffraction experiments while analysing shifts in basal spacing reflections [183]. Smectite and CO2 interaction has also been investigated with neutron diffraction and infrared measurements [184,185,186]. Smectites are layered nanoporous materials, and it was found that they can sorb CO2 in their interlayer space due to gaseous CO2 intercalation into the interlayer nano-space as confirmed using above mentioned techniques [28]. In addition, experiments with smectites have shown that Na-exchanged montmorillonite can swell up to 9% if it is exposed to high purity CO2, this however depends on the initial hydration state of the interlayers in smectite [24]. CO2 adsorption within the smectite interlayer region in Na-exchanged montmorillonite is believed to be the reason behind clay swelling up to the accounted 9% [24]. Moreover, recent studies have shown that the pore size, charge, and solvation energies of the interlayer cations in smectite minerals are greatly affecting the structure, dynamics, and energetics of the intercalated CO2 molecules [187,188]. At the same time, other studies have shown that, for instance, illite and kaolinite minerals are not known for interlayer expansion under any experimental conditions, whilst they are still able to adsorb considerable amounts of CO2. It is believed that in these clay minerals, CO2 sorption occurs on clay platelet surface [186]. Accordingly, it can be concluded that CO2 adsorption mechanisms vary depending on clay mineral type as it can be either CO2 adsorption in the interlayers, or it can be surface sorption. These regularities must be further addressed in future experiments.
One of the main factors affecting the physical sorption of CO2 onto clay minerals is the type of exchangeable ion within; for example, Cs-exchanged and tetramethylammonium-exchanged montmorillonite have sorption capacity for CO2 up to 1.70 mmol g−1 even at ambient pressure and temperature [189]. At the same time, in case of fibrous clay minerals, such as sepiolite, the sorption takes place due to the presence of nanocavities, which are acting as a molecular sieve and therefore the CO2 uptake can reach up to 1.48 mmol g−1 [119]. In comparison, in saponite (smectite group), CO2 physical sorption takes place under ambient conditions on the nanosheet surface at the open spaces on clay mineral [190]. Additionally, a number of other clay minerals have also been studied as potential CO2 adsorbents, for example, hydrotalcite and hectorite, palygorskite, kerolite, and stevensite, thus demonstrating the high potential of clay minerals for carbon capture [29,119,191,192,193].

7. Prospective for Clay Mineral Modification to Improve Carbon Dioxide Capture

Physicochemical characteristics of clay minerals have been widely studied using numerous techniques and approaches, such as neutron and X-ray scattering, nuclear magnetic resonance (NMR) spectroscopy, tracer experiments, numerical modelling, and others [25,99,155,156,194,195,196,197,198,199,200,201]. According to the data obtained from using these approaches, clay mineral properties, in regard to the improvement of CO2 adsorption, can be optimized by the insertion of organic, inorganic, or organometallic species between adjacent sheets of clay structure, by functionalization or by acid treatment [27,138,139,193,202]. However, before choosing one or another type of clay mineral for carbon capture, it is important to note that the chemical properties of particular minerals, and consequently also the adsorption of gases on or within them, are significantly affected by the structure of the mineral and the charge of clay layers. Mineral structures with all six octahedral sites occupied are known as trioctahedral structures, while structures with four out of six octahedral sites occupied are referred as dioctahedral structures. For example, minerals such as kaolinite, halloysite (1:1 layer structure), and montmorillonite (2:1 layer structure) are dioctahedral, while serpentine (1:1 layer structure) and vermiculite (2:1 layer structure) are trioctahedral. Furthermore, structures of palygorskite and sepiolite are rather different from above mentioned due to the lack of continuous octahedral sheets; however, there are fibrous clay minerals with dioctahedral and trioctahedral arrangement [203]. At the same time, the structure of a layer, its thickness, and the charge are all interrelated with one another and can therefore affect the particle size and specific surface area of the clay mineral and many other properties [204]. In addition, recent studies indicate that raw clay materials typically show low CO2 adsorption capacity, and the adsorption properties are generally related to the textural properties of particular minerals [119,121,205,206]. Therefore, clay mineral modification to improve CO2 capture is a crucial direction of the future research [207,208].
Factors/parameters that are essential for the selection of a potential sorbent material include high CO2 absorption capacity, high CO2 selectivity, mechanical, hydrothermal, and chemical stability, rapid adsorption/desorption kinetics, regeneration capacity, low heat capacity, price of raw materials, and costs of synthesis [99,121,209,210]. Therefore, appropriate materials are chosen either based on the highest efficiency within these parameters or materials are modified to improve certain parameters. At the same time, it is important to note that it is almost impossible for a single material to satisfy all of the above-mentioned essential properties of a sorbent. Nonetheless, it is important to assess at least partial compliance and to consider the most important factors.
Efficiency-wise, clay minerals are among the materials with the most essential properties required for CO2 sorption. In addition, porous clay heterostructures have been evaluated as adsorbents for small hydrocarbons, in their separation and in the encapsulation of organic compounds [192,211,212,213]. There have also been experiments on clay structures as the catalytic support in the NO reduction, partial oxidation of H2S, preferential oxidation of CO2, and the esterification reactions to obtain bio-lubricants and all of these experiments have shown high potential of clay applications [214,215,216,217].
Strategies to overcome drawbacks for CO2 sequestration encourages capturing and storing CO2 using solid adsorbents, which operate through a repetitive adsorption–desorption cycle, thus using sorbents that display base properties [218]. Solid sorbents usually contain a solid with a large specific surface area, and when a layer of such sorbent passes through a mixture of gases, CO2 is captured by either a physical sorption or a chemosorption process. These sorbents can be regenerated via pressure or temperature change [219,220,221]. For example, pressure changes have been successfully used to induce adsorption–desorption cycle and to separate CO2 from N2 at room temperature. In this regard, montmorillonite clay intercalated with dendrimer polyol silica, goes through a pressure-induced adsorption-desorption cycle and shows a potential for reversible CO2 capture. Moreover, the hydroxyl groups in montmorillonite are found to act as adsorption sites and interact with CO2 molecules in the gas phase [29,170,174,218,222,223,224,225,226,227]. Numerous clay mineral modifications, such as clay-graphene hybrids, have also been assessed for CO2 adsorption [228]. Clay minerals are considered suitable for further modifications involving a simultaneous reduction of basicity and an increase in CO2 sequestration [218].
One of the options for clay mineral modification to improve the CO2 sorption capacity is their expansion. For instance, smectites are expanded by incorporation of bulky cations in the interlayer space [229]. Additionally, these modified clays are with significantly improved thermal and mechanical resistance when compared to a raw smectite [229,230]. The reactivity of smectite group minerals can be further improved by replacing mobile and exchangeable cations on their surface by higher valence metal ions (e.g., Mg2+, Zn2+, Cu2+, Al3+ or Fe3+) [231,232,233,234]. Furthermore, the cation exchange with Ni2+ ions can promote formation of a carbonaceous residue [235,236,237]. It can be concluded that amongst the most significant properties of clay minerals to support CO2 adsorption is the ion exchange. The properties of ion exchange allow for isomorphous substitution of metal cations in the lattice by lower-valent ions. The created negative charge can then be balanced by other cations. The cations can be exchanged when they are brought into the contact with other ions in aqueous solution [137]. Considering cation exchange properties and layered structure of clay minerals, it is possible to introduce molecules of other compounds, which then can be used to change and modify clay mineral properties.

8. Clay Mineral Activation

Nowadays, a wide variety of clay modification approaches have been used to improve CO2 sorption capacity. In addition, various modification combinations to extend the application range of these materials have been approached. The modification typically is carried out using either raw clay material or activated clay material. The textural and chemical properties of clay materials can be changed essentially by clay activation. The most commonly used clay activation methods are mechanochemical activation, intercalation, thermochemical activation, and chemical activation. All these approaches have high potential to promote the generation of new clay pore structures and formation of new or improved active sites on the surface of clay materials [132].
The most common method for clay activation is the activation by alkaline or acid treatment (chemical activation). For example, Wang et al. (2013) proposed to improve kaolinite and montmorillonite textural properties through acid or alkaline treatment, and it was found that the acid treatment affected the clay structure to a higher extent than the alkaline treatment [132]. In general, the acid activation is a chemical treatment to alter clay textural (surface area, pore volume) and chemical properties (catalytic, adsorptive etc.). In addition, the acid activation also includes clay leaching with inorganic acids, causing disaggregation of clay particles, elimination of mineral impurities and dissolution of external layers [132]. Additionally, Horri et al. (2019) have studied the impact of acid treatment time (3, 8, and 24 h) using bentonite and 3 M HCl, and it was found that the content of Al decreased significantly with increasing treatment time, while the percentage of Fe and Mg oxide significantly decreased only after 24 h. Moreover, the specific surface area of 3 h treated bentonite was 227 m2 g−1, which was significantly higher than that for raw bentonite (21 m2 g−1). It is believed that the acid treatment leads to an increase in the specific surface area and pore volume because of the pore opening that appears due to the octahedral sheet degradation. However, longer acid treatment times reduce the specific surface area, as in the study performed by Horri et al. (2019). After 8 h, the specific surface area reduced to 223 m2 g−1 and 177 m2 g−1 after 24 h, respectively [206]. Even if there is no clear proportionality between sorbed CO2 amount and the acid treatment time, it can be observed that the acid treatment significantly increases the sorption capacity of the material when compared to a raw bentonite [206]. At the same time, it is believed that the acid concentration plays an important role in the activation process. For example, Wang et al. (2013) reported that when using either 3 M, 6 M, or 9 M H2SO4 for the activation of bentonite, the use of 6 M H2SO4 gave the most optimal results, while more concentrated acid solutions led to pore collapse, which ended up with a smaller surface area and pore volume [132].
The efficiency of acid treatment can be optimized using such an approach as the microwave-assisted radiation, because it is possible to increase the specific surface area of the clay minerals using shorter treatment time and more diluted acid solution [99,132]. For instance, Cecilia et al. (2018) confirmed that the microwave-assisted acid treatment for sepiolite and palygorskite has resulted in an increased surface area and pore volume of clay minerals [119]. Similar results were also obtained by Franco et al. (2020) when studying microwave-assisted acid treatment of kerolitic clay, where after the acid treatment the surface area varied from 250 to 435 cm2 g−1 and the CO2 adsorption increased to a greater or lesser extent, depending on a particular clay material that was used [193]. In addition, Cecilia et al. (2018) reported that the acid treatment reduced the CO2 adsorption capacity for sepiolite but increased it for palygorskite. These findings coincide with other authors’ conclusions that the treatment with acid is causing the loss of cations in the interlayer distance and cations in the octahedral sheet [99,119]. It is worth noting that even though the acid treatment is increasing clay textural properties (e.g., pore size, pore volume); however, this process is not always beneficial to improve the CO2 adsorption capacity. For instance, Irani et al. (2015) have studied sepiolite treatment with 2 M HCl and described the obtained material using energy dispersive and infrared spectra, which have shown that during the acid treatment Mg2+ ions were completely extracted from the sepiolite crystalline lattice, whereas Si-OH had formed in the structure of treated sepiolite. The surface area for raw sepiolite used for their experiments was 103.7 m2 g−1, but after the acid treatment it significantly increased and reached 272.4 m2 g−1 [177]. In yet another study, Cecilia et al. (2018) have reported that the microwave-assisted acid treatment had increased the surface area from 182 m2 g−1 to 326 m2 g−1, while CO2 adsorption reduced from 65 mg g−1 to 41 mg g−1 at 25 °C and 1 bar. The possible reason for this is related to the partial collapse of the octahedral layer, which may cause an increase in the size of the nanocavities and thus resemble something similar to molecular sieve that retains CO2 molecules due to a weaker physical interaction between CO2 molecules and the wider nanochannels of the acid treated clays. Other possible reason can also be related to loss of cations in the nanocavities that interact with CO2 molecules [119].
A notable approach for clay modification is the creation of pillared interlayered clays (PILCs). It is a class of two-dimensional microporous and synthetic materials, and due to the high surface area and permanent porosity of PILCs, these are very attractive solids for CO2 capture. In this regard, smectite group is the most preferred for modification. Minerals from the smectite group are layered clays and the charge is balanced between their layers by interlayer cations, such as Na+ or K+. Exchanging these cations for other inorganic species allows to construct PILCs. The exchange is often performed for large oligomeric polycations, such as Zr, Al, Fe, Cr, and various other oligomers [238,239].

9. Clay Mineral Functionalization

Besides the activation, there are several other appropriate methods for clay modification, among which the functionalization by grafting and/or the impregnation (immobilization) with amine containing polymers are showing considerably high efficiency. For instance, the functionalization of clay minerals with amine groups improves the CO2 adsorption capacity due to the synergetic effect of the molecular sieve of the adsorbent and the chemical interaction between the amine species and CO2 molecules [90,240]. The functionalization is performed by using the grafting and impregnation of amine-rich polymers. The grafting is a chemical process, where silanol groups located on the surface of the adsorbent react with amino alkoxysilane molecules to form a material with available amine groups on its surface [94,122]. The impregnation of amine-rich polymers is stabilized by hydrogen bonds with the silanol groups of the adsorbent [95,123,241].
Clay minerals can be substantially modified by replacing the natural inorganic interlayer cations with selected organic cations [137]. An example of organic compounds used for intercalation are amines and amino acids [242,243]. Since amines can react selectively with CO2, various attempts have been made to incorporate amines into porous materials. These materials are chemically treated with amine containing compounds, thus obtaining new sorbents. When a layer of such a sorbent passes through a mixture of gases containing CO2, the immobilized amine groups react with CO2 to form carbamates, resulting in the capture of CO2. Amine-containing sorbents are receiving increasing attention due to such advantages as low energy consumption, high CO2 capacity, high resistance against contaminants, and high stability. Furthermore, recent research indicates that amine efficiency increases with increasing amine loading onto the sorbent that consequently provides much stronger CO2 binding to a high density of amine loading. It is also important to underline that the adsorption efficiency is affected by the reaction conditions, either dry or humid conditions [244]. One of the proposed mechanisms is that under humid conditions water or hydroxide ions can act as a base and the amine efficiencies can approach the unity, while another proposed mechanism explains that under humid conditions ammonium bicarbonate species are formed [99,244]. In turn, under dry conditions a second amine acts as a base to produce an ammonium carbamate, giving a theoretical maximum efficiency of 0.5. Thus, obtaining materials with high density of amines it is possible to improve the efficiency of CO2 capture [99,206].
The functionalization of porous materials with amine species (primary or secondary) improves the chemical interactions between the given adsorbent and the CO2 molecules via zwitterion intermediate to form ammonium carbamate species [72,99]. The carbamate is formed by a nucleophilic reaction between the lone electron pair on the nitrogen of the amine and CO2, during the anhydrous adsorption process. This process results in the formation of a zwitterion (equal number of positively and negatively charged functional groups), which is deprotonated in the presence of a base [206,244,245]. The tertiary amines interact with the CO2 and produce bicarbonate [99]. Three main clay mineral functionalization methods involve:
  • Impregnation of amines onto porous carriers;
  • Formation of covalent bonds between amine containing functional groups and porous surface (grafting);
  • In-situ polymerization.
Wet impregnation can be used to immobilize amines into pores if the solid support material is porous. Additionally, the amine compounds can be covalently bound to solid support material via hydroxyl functionality. Another approach is a direct polymerization of amine containing polymers on the surface of solid support [210]. In this regard, solid support materials with a high specific surface area and a porous structure are commonly used.
Impregnation of porous adsorbent with amine containing polymer is a widely used method to disperse a high amount of amine species on the surface of the adsorbent. Yuan et al. (2018) suggest that wet impregnation facilitates much higher loading of accessible amine molecules than is achievable by grafting or co-condensation, and this is resulting in an enhanced CO2 adsorption capacity [178]. The most often used wet impregnation reagents are polyethyleneimine (PEI), tetraethylenepentamine (TEPA), triethylenetetramine (TETA), diethanolamine (DEA), diisopropanolamine, triethanolamine, and diethylenetriamine (DETA) [206,246]. In addition, Chen and Lu (2014) have studied kaolinite wet impregnation using monoethanolamine (MEA), ethylenediamine (EDA), and a mixture of both (4MEA+1EDA) and found that the latter shows the highest CO2 sorption capacity [205]. Furthermore, Atilhan et al. (2016) have modified montmorillonite nanoclays with various amino groups and revealed that the impregnation of montmorillonite nanoclays with a single primary amine (octadecylamine) is more effective than the impregnation with tertiary amine (dimethyl dialkyl) or modification with doubling primary amines (octadecylamine) [247]. It is believed that the modification with primary amines is enhancing the hydrophilic character of nanoclays and thus is also affecting the amount of sorption sites. In similar studies, Ouyang et al. (2018) studied the impregnation possibilities of the acid treated sepiolite using five different amines, from which ethylenediamine was the only containing terminal –NH2 groups, while TETA, TEPA, and PEI contained both terminal (-NH2) and middle (-NH-) radicals and the length of their chain increased gradually until up to polymers with random networks [248]. The highest CO2 adsorption capacity (up to 3.7 mmol g−1 at 75 °C, 60 mL min−1 CO2, and 40 mL min−1 N2 mixture) was obtained using the 50% TEPA-loaded modified sepiolite matrix, when compared to other materials using the same concentration of amines. The authors revealed that the sorption capacity for materials containing other amines were 2.48 mmol g−1 for PEI, 1.99 for TETA, and 1.68 mmol g−1 for ethylenediamine, respectively [248].
Several authors have explored the impregnation with amine containing polymers at different loadings [107,178,205,209,246,249]. From their research results, it can be concluded that the CO2 adsorption capacity increases with an increase in amine loading, and this is due to more active amine sites in the surface layer of the sorbent; however, it decreases if there is a blockage in the mesopores. For instance, Wang et al. (2014) have studied various PEI loadings on the acid-treated montmorillonite and found that the most optimal PEI loading amount was ≤50 wt%, which reached the highest CO2 sorption capacity (112 mg g−1) at 75 °C under dry conditions, while further enhancement up to 142 mg g−1 was observed with the addition of moisture [107]. The authors have found that the CO2 sorption capacity decreased when PEI loadings were further increased, which was possibly due to fast decrease of the accessible amine sites, possibly due to the agglomeration of the particle. Similar results were also obtained by Niu et al. (2016), who used halloysite nanotubes that were pre-treated to produce mesoporous silica nanotubes, which were further impregnated with PEI, where PEI loadings varied from 30 to 60 wt%. It was found that 50 wt% loading of the PEI was the most suitable and the sorbed CO2 amount reached 2.75 mmol g−1 at 85 °C in 2 h [209]. Furthermore, Ouyang et al. (2018) reported that MgO-SiO2 nanofibers from sepiolite loaded with 50 wt% of PEI reached higher CO2 sorption capacity (2.48 mmol g−1 at 75 °C) than using 30 wt% to 60 wt% PEI loadings [249]. Analogous results were also obtained by Zhang et al. (2020) while studying CO2 adsorption on an exfoliated vermiculite-TEPA composite, where TEPA content in the composite (0.2 to 50 wt%) strongly affected the textural properties of the composite and sequentially also the sorbed amount of CO2, indicating that the optimal TEPA loading was 2 wt% [246]. Meanwhile, Wang et al. (2013) explored the wet impregnation of bentonite using varied loadings of TEPA [132]. The highest adsorption capacity was obtained for the material modified with 6M H2SO4 and impregnated with 50 wt% of TEPA at 75 °C, while loading higher than 50 wt% had negative impact on the adsorption performance. Irani et al. (2015) have studied the effect of TEPA loading on the acid treated sepiolite and found that the increase of TEPA loading from 30 to 60 wt% increased CO2 capacity from 1 to 3.8 mmol g−1. At the same time, further increase of TEPA loading from 60 to 70 wt% decreased CO2 capacity significantly [177]. Liu et al. (2018) have explored CO2 sorption onto purified, acid-treated sepiolite impregnated with various amounts of DETA and noticed that sepiolite loaded with 0 to 0.2 DETA has lower CO2 capacity in comparison to acid-treated sepiolite, which was possibly due to the reduction of physisorption because of the blockage of pores with amine species [250]. Further increase of DETA loadings (up to 1) led to reduction of adsorbed CO2 and this finding coincides with previously mentioned results obtained by other authors using PEI and TEPA loadings. Yuan et al. (2018) have studied CO2 adsorption (50 °C in a pure CO2 atmosphere with a flow rate of 100 mL min−1) using TETA loaded acid treated sepiolite, where TETA loading varied from 10% to 60% [178]. It was found that loading with TETA significantly increased sorption capacity of sepiolite. Moreover, TETA loading increased in sepiolite from 10% to 30% and the resulted increase of sorbed amount of CO2 was from 1.28 to 1.93 mmol g−1, accordingly. Further increase of TETA loading in sepiolite (30% to 60%), reduced sorbed amount of CO2 from 1.93 mmol g−1 to 0.74 mmol g−1.
Another functionalization method, grafting, involves a chemical reaction between the available silanol groups of the surface of the adsorbent and amine alkoxysilane compound to obtain hybrid adsorbents with high thermal stability, high water tolerance, and high selectivity towards CO2 [206].
The amine-based materials obtained by the interaction, in which the amine species are incorporated in the surface of the adsorbent, are characterized by high thermochemical stability. Amino-containing organosilanes containing moieties such as amino-propyl (AP), ethylene-diamine (ED), diethylene-triamine (DT), 3-aminopropyl-trimethoxysilane (APTES), N1-(3-trimethoxysilylpropyl) diethylenetriamine (TMSPDEA), tetraethylenepentamine (TEPA), and N-2-aminoethyl-3-aminopropyltrimethoxysilane (AEAPTS) are often used as the grafted compounds.
Grafting involves interaction between amine species and CO2 molecules via zwitterion mechanism that results in the formation of carbamates [206]. A number of authors have researched the sorption of CO2 on grafted clay minerals. A variety of clays and clay-containing materials have been used, such as bentonite, saponite, palygorskite, montmorillonite, sepiolite, and kaolinite. Stevens et al. (2013) have studied the introduction of amine groups on the surface of montmorillonite or hexadecyltrimethylammonium bromide-intercalated montmorillonite (MMT CTAB N2) using a water-aided exfoliation method [251]. Within this approach, the desired amount of material was inserted in toluene and the reaction mixture was then sonicated in an ultrasound bath for 4.5 h at room temperature. The authors explained that the sonication was used to delaminate larger particles and thus to significantly increase the specific surface area and consequently to reach much higher amine grafting rate and less pore blockage. Afterwards, AEAPTS was added to the reaction mixture and sonicated for another 24 h at 60 °C. Samples were then filtered and dried at 80 °C overnight. The authors revealed that the highest CO2 adsorption capacity (2.4 mmol g−1 at 100 °C) was reached using initially intercalated and then grafted montmorillonite (MMT CTAB N2), and they explained that chemosorption is the predominating process there. Furthermore, such sorbent showed rather good stability in pure CO2, while its CO2 sorption ability was significantly reduced in the presence of SO2 [251]. Simultaneously, Gomez-Pozuelo et al. (2019) reported that the adsorption capacity of CO2 for bentonite, montmorillonite, saponite, palygorskite, and sepiolite functionalized by grafting with the (3-aminopropyl)-trimethoxysilane and N1-(3-trimethoxysilylpropyl) diethylenetriamine ranged from 32 mg g−1 to 61 mg g−1 at 45 °C under 1 bar [121]. In addition, Horri et al. (2019) have studied CO2 adsorption abilities on APTES and diethylenetriamine-trimethoxysilane (DT) grafted bentonite treated with 3M HCl for 3, 8, and 24 h before the functionalization [206]. The results have shown that grafting with APTES increased the sorption capacity of the material, while the use of DT reduced the overall adsorption capacity. Different sorption capacities of the materials after their functionalization with either APTES or DT can be explained by the longer DT chain length, which may result in the pore blockage and thus a significant part of the amino groups may not be accessible to CO2 molecules. The authors explained that the material may have been saturated with DT. The solution to overcome this issue could be grafting with DT using another support material with a larger specific surface area and pore volume, avoiding the saturation of the material [206].
Vilarrasa-Garcia et al. (2017) have studied the incorporation of amine species on bentonite and porous clay heterostructures (PCH) using grafting by APTES or impregnating by PEI or TEPA [74]. The authors revealed that PCH is the most appropriate material for the functionalization. In addition, CO2 sorption on grafted and impregnated materials occurred via chemical and physical interaction, due to co-existence of these dual sites [74]. Furthermore, functionalization with APTES indicated the contribution of both sites, while functionalization with PEI and TEPA favoured the chemical contribution, especially as the polymer content in the material increased. As the polymer content in the material increased further, CO2 adsorption capacity decreased; this is due to stacking of the amine-rich polymer. Similar results were also reported by Cecilia et al. (2018) by showing that the grafting with APTES or impregnation with PEI is decreasing CO2 adsorption, while it increases the chemisorbed amount of CO2 [119].
Several authors have offered to use the aminosilane-modified clay nanotubes to entrap CO2 [214,244]. Following this approach, three different aminosilanes—(3-aminopropyl)triethoxysilane, N-[3-(trimethoxysilyl)propyl]-ethylenediamine, and N1-(3-trimethoxysilylpropyl)diethylene-triamine have been used [244]. The highest CO2 adsorption capacity was reached in halloysite nanotubes modified with N1-(3-trimethoxysilylpropyl)-diethylenetriamine; while CO2 adsorption capacity of halloysite nanotubes without amine loading was extremely low (almost non-existent) even at the same experimental conditions [244]. In addition, Jana et al. (2015) have studied the isotopic selectivity of CO2 adsorption on amine grafted halloysite nanotubes, studying major abundant isotopes of CO2 [245]. This experiment was carried out using an optical cavity-enhanced integrated cavity output spectroscopy. Results suggest that amine-grafted halloysite nanotubes can be regenerated at relatively low temperatures, and thus recycled repeatedly to capture atmospheric CO2 [245]. Cecilia et al. (2018) reported that the acid treatment improves the functionalization by grafting due to a slight increase of the pore size resulting from partial or total pore blockage of the nanochannels of the fibrous clay materials [119].
It is suggested to incorporate amine species into the clay minerals by a double functionalization method. This approach involves grafting and impregnation procedures consequently on the same material [119,121]. The limited surface groups or adsorption sites may restrict the number of grafted amine groups, whereas the combination of chemical grafting and physical impregnation can incorporate unlimited amine groups onto the carrier. Gomez-Pozuelo et al. (2019) have noticed that as a result of organic functionalization of clays, the texture properties of grafted and impregnated clays decrease [121]. The authors presented the decrease of the surface area from 137 m2 g−1 for unmodified palygorskite to 55 m2 g−1 for grafted clay samples and 42 m2 g−1 for impregnated clay materials. Similarly, the pore volume decreased from 0.32 cm3 g−1 to 0.16 cm3 g−1 and 0.11 cm3 g−1 for grafted and impregnated materials, accordingly. The authors have noticed that amino-containing organosilanes that contain more amino groups also have higher CO2 sorption capacity. Furthermore, PEI impregnation and double functionalization yielded higher organic loadings but lower amino efficiencies than grafted samples. At the same time, the double functionalization showed lower CO2 adsorption than individual grafting or impregnation due to pore blocking by high organic loadings. Simultaneously, the sorption efficiency was higher in humid environment and grafted materials were more stable than PEI-impregnated materials after three adsorption/desorption cycles [121].
Cecilia et al. (2018) suggested an increase of available amine sites or higher proportion of primary amines obtained after grafting with APTES [119]. In their study, the CO2 uptake of sepiolite after double functionalization reached 1.41 mmol g−1 at 760 mmHg and 25 °C, while the raw palygorskite reached 1.04 mmol g−1 under the same conditions. The double functionalization (APTES-PEI) led to the highest adsorption capacity due to the higher amount of available amine sites that favour the chemical interaction with CO2 molecules [119]. Roth et al. (2013) have studied double functionalization of montmorillonite nanoclays using at first grafting with 3-aminopropyltrimethoxysilane (APTMS) and afterwards, the wet impregnation of PEI, to achieve 50% loading of PEI-treated clay. The authors reported that the double functionalization of montmorillonite nanoclays allowed to improve the sorption capacity of CO2 up to 7.5% at 85 °C at the atmospheric pressure [210].
The affinity of untreated clays for CO2 is enhanced by the intercalation of organic matter with compounds with basic properties, such as amine compounds, polyol, and amino dendrimers with hydroxyl and amino groups, respectively [252]. By intercalating clays with polyol dendrimers, it is possible to obtain organoclays, which recently have been used also for CO2 capture [99,218]. Some authors suggest the intercalation of polyols-species in the interlayer spacing, thus it would be possible for CO2 molecules to interact with weak base -OH sites. Some reports indicate that montmorillonite intercalated with polyols dendrimers shows slightly lower affinity to CO2, while the regeneration is easier in comparison to raw support material. Thus, this indicates that weak basicity can promote the reverse capture of CO2. It is suggested that an ideal adsorbent would be able to release the adsorbed gas upon slight heating or under vacuum [218]. Furthermore, Azzouz et al. (2013) have explored the role of -OH groups when preparing polyol-montmorillonite using different polyalchohols and found that the incorporation of polyalcohol molecules significantly enhanced the affinity of montmorillonite towards CO2 and the OH groups of the incorporated polyalcohol were the main adsorption sites [218].
There were several attempts to insert a cationic amine-rich dendrimer in clay material, for example, inserting polyamidoamine (PAMAM) to improve CO2 sorption capacity [99]. The organo-clays present are three adsorption sites; two of them are attributed to the clay layers (internal binding unit and external binding unit) and the third one is the adsorption site attributed to the availability of the dendrimer sites, which grows directly with the amount of intercalated-dendrimer. This site is the most determining since the CO2 adsorption capacity increases directly with the amount of the intercalated dendrimer as the consequence of the strong affinity between the dendrimer and CO2 molecules [99]. Stevens et al. (2013) have prepared hexadecyltrimethylammonium bromide (CTAB) intercalated montmorillonite and used it as a source material for further amine surface modification, as well as studied CO2 adsorption process using it as adsorbent [251]. In this experiment, the highest CO2 sorption capacity (0.19 mmol g−1) was reached at 25 °C and the capacity was decreasing with increasing temperature, thus indicating that the predominant process is physisorption.
Pires et al. (2018) offered the amino acid-intercalated montmorillonite as an alternative sorbent material for CO2 sorption. The main advantage of such a material is environmentally friendly synthesis that can be performed in water environment without using noxious solvents. It also has relatively lower raw material costs and therefore no expensive amine alkoxysilanes, instead using renewable non-toxic reactants, as well as easy preparation [253]. The authors concluded that amino acid intercalated montmorillonite promotes the sorption of CO2 in comparison to raw montmorillonite. Furthermore, the number of amino groups per molecule that can interact with CO2 had a major impact on the ability to capture the gas [253]. Simultaneously, Elkhalifah et al. (2015) offered amine-bentonite hybrid sorbent for CO2 adsorption. The preparation of proposed sorbent involves two steps—preparation of magnesium form of bentonite and sorbent synthetization by intercalating monoethanolammonium cations in the interlayer space of the Mg-bentonite [252].
The sorption of CO2, for instance, on amine-containing porous sorbents includes both chemosorption on amino groups and physical adsorption within pores. Therefore, the adsorption temperature has a significant effect on which of the sorption modes will be predominant should be accounted for when modifying clay minerals. Wang et al. (2014) have studied the impact of temperature and found that with increasing temperature, the sorption capacity of montmorillonite/PEI sorbent increases up to 75 °C but decreases from 75 to 85 °C. It is suggested that the CO2 diffusion is the predominant at the temperature range from 30 to 75 °C, but above 75 °C the thermodynamics become dominant, and the equilibrium shifts to the desorption, resulting in the decrease of CO2 sorption capacity [107,132]. Similar results were obtained by Chen et al. (2013) when studying CO2 sorption on PEI-impregnated bentonite at the temperature interval 25 to 100 °C and found that 75 °C is the optimum temperature for CO2 capture. The authors explained that the temperature increase promotes a flexibility of amine groups and thus increases the affinity of these sites to CO2, at the same time, too high temperature may also cause the desorption of CO2 [254]. Cecilia et al. (2018) confirm that CO2 adsorption capacity improves with increasing temperature thanks to a rearrangement of the amine-rich polymer favouring the diffusion of the CO2 molecules in the adsorbent, therefore enhancing the CO2/N efficiency [119]. Ouyang et al. (2018) also reported that PEI-loaded MgO-SiO2 nanofibers from sepiolite reached the highest adsorption capacity of CO2 at 75 °C [248,249]. Similar conclusions were reached by Irani et al. (2015) in studying the CO2 sorption capacity of TEPA-impregnated sepiolite in the temperature range of 25 to 70 °C. As the temperature increased from 25 to 60 °C, the sorption capacity increased, but it decreased with increase of temperature from 60 to 70 °C [177]. Yuan et al. (2018) found that the CO2 adsorption capacity on TETA-impregnated sepiolite decreased with increasing temperature in the interval from 30 to 70 °C [178].
In-situ polymerization requires placing a monomer between clay mineral layers and the expanding and dispersing the mineral layers into the matrix by polymerization. This approach can significantly increase the surface area and consequently also improve the CO2 adsorption capacity. At the same time, the exact properties of clay-polymer composite are largely affected by polymer-polymer and polymer-clay interaction, as well as clay aspect ratio, dispersion, and the alignment, all of which are highly variable characteristics among clay minerals [255,256,257].

10. Clay Mineral Assemblages in the Baltic States

The Baltic region, or Baltic Sea region, refers to countries in the general area surrounding the Baltic Sea. Countries with shorelines along the Baltic Sea include Denmark, Estonia, Latvia, Finland, Germany, Lithuania, Poland, Russia, and Sweden. In this review, authors focus on the Baltic States to explore the potential of sustainable and cost-effective local natural resource usage of the region, conjointly with the GHG emission reduction in these post-soviet countries. The Baltic States is a geopolitical term used to group three countries on the eastern coast of the Baltic Sea: Estonia, Latvia, and Lithuania. All three countries are members of EU and engage in the implementation of the European Green Deal.
The Baltic States are located on the Eastern European Craton, which is the core of the Baltica proto-plate. Three countries are located on the Baltic Shield of the Eastern European Craton. The Baltic Shield is the exposed Precambrian northwest segment of the East European Craton. The Baltic Shield contains the oldest rocks of the European continent with a lithospheric thickness of 250 to 300 km [258,259]. Generally, in the direction from the northeast to southwest throughout the Baltic States, the stratigraphical layers of the craton are monoclinally deepening from the north of terrestrial Estonia to Lithuania, where the craton reaches a depth of more than 2 km.
Under the Quaternary sediment bedrock surface in the Baltic States, deposits of Cambrian, Ordovician, Silurian, and Devonian strata (in Estonia); mainly Middle and Upper Devonian (in Latvia); in the south-eastern part of Latvia and Lithuanian territory also Carboniferous, Permian, Triassic, Jurassic, and Cretaceous (only in Lithuania) are represented [260].
Due to the geological history of the region, clay in the Baltic States is among the most common types of deposits. During the Quaternary deglaciation of the region, glaciolimnic sediments on the surface were deposited in the form of varves into waterbodies in front of glacier margins [261]. In addition, due to the wide distribution of clay deposits on the surface of Quaternary environment, large territories have undergone paludification and transformed into mires [261]. Simultaneously, in smaller quantities also Devonian, Jurassic, Triassic, and even Cambrian clays are present in the region.
Use of clays in the Baltic region has been known for thousands of years, for example, at about 6000 years before present inhabitants of Estonia learned to make earthenware from locally available clay deposits. As civilisation has evolved, so has the pattern of clay use. Locally available clays were used in the production of worldwide recognizable red bricks, which were used as building material for strongholds and churches and provide modern Estonia’s historical buildings and architectural monuments with distinct geological grandeur [261]. A major amount of these clays is of Cambrian origin. Lower Cambrian (Lontova formation) clays in north Estonia are represented at shallow depths and contain rare earth element deposits. Clays of Lontova formation can also be found in the outcrops of the Baltic Klint [262].
Clay deposits in Latvia generally are related to Devonian and Quaternary systems, although rare deposits of Triassic and Jurassic clays are also present here. The most recognizable clay in Latvia is Liepa (Lode) clay from the Devonian period, Saltiski clay from the Triassic period, and Apriki clay from Quaternary period. These clays are dominated by illite minerals with additional kaolinite minerals in varied proportions [263]. Devonian clay is distributed in the north and northeast Latvia and can be extracted from numerous mineral deposits, however, the only mineral deposit currently active for extraction is Liepa (Lode) clay deposit (Cesis municipally) [264]. Devonian clays contain a high percentage of illite minerals, with additional kaolinite and chlorite minerals. Triassic clay is rare and can be found only in few mineral deposits in the southeast Latvia. Triassic clays contain high percentage of montmorillonite minerals. Triassic clays have high adsorption capacity. A large capacity of Triassic clay is in Lithuania; however, they are industrially exploited. Jurassic clay is rare and can be found only in close proximity to Triassic clay deposits in the southeast Latvia. One of the major distinctive features of Jurassic clays is the presence of natural organic matter within these deposits [265]. At the same time, in Lithuania parts of Jurassic clay deposits are exposed to the surface (Papile region), while in Estonia these clays are not present. The origin of these black clays that are rich in organic matter is still under dispute; perhaps they were deposited in shallow lagoons where large rivers transported organic sediments during Mesozoic era [265]. Quaternary clays can be found in mineral deposits throughout Latvia and in less amount in Lithuania and Estonia. These are typical glaciolimnic clays.
It is important to emphasize that the exact amount of clay stocks in the Baltic region is not known. This is due to diverse levels of research work on clay deposits, which therefore cannot be mutually comparable. Moreover, the majority of clay deposits do not have their stocks recalculated to clay reserves. For instance, while the quantity of clays from Lontova formations (Lower Cambrian) are almost precisely known, the quantities of Quaternary glacial clays are often only generalized calculations that were performed during the USSR period. At the same time, the amount of clay stocks is not the key factor for the evaluation of clays for CO2 capture. In turn, adsorption/absorption parameters of clays play the key role here.

11. Discussion

Major climate change is inevitable. However, if action is taken now, then the rapid rise of temperature could be slowed down and the impact on the environment could be less damaging. The climate crisis demands immediate action towards carbon neutrality and carbon negative solutions. Unfortunately, the climate change will also necessitate adaptation to the new climatic conditions. Reducing emissions at this point is critical so that the adaption is gradual.
One of the options to mitigate CO2 emissions is their capture from emission sources, such as industrial objects. Even though there are numerous media with a potential to capture CO2, clays or any other amorphous solids seem to be amongst the most reasonable options when considering efficiency, price, maintenance, disposal, and other aspects. In this respect, clays are among the lowest-cost materials. At the same time, it must be considered that different clay minerals have rather different abilities to capture CO2. For example, research shows that montmorillonite clay mineral has the highest CO2 adsorption capacity, followed by illite and kaolinite [266]. Notably, that variation in the CO2 adsorption capacity among different clay minerals is also related to cation exchange capacity [267].
Montmorillonite is the most common smectite group mineral; it is built of several tetrahedral silicate sheets with one octahedral metal oxide or hydroxide sheet in between them. This composition allows montmorillonite to possess large surface area, swelling abilities, numerous functional groups, and ion exchangeability, all of which contribute to the high CO2 sorption efficiency [139,268,269]. Montmorillonite is composed of octahedral sheets of aluminium or magnesium oxide, which is located between tetrahedral silicone layers. While oxide sheets are negatively charged, silicone layers are positively charged. The charges are balanced by the counter cations located in between the interlayer space [270,271]. The highly porous nature of montmorillonite and substantial number of cations within the structure and high cation exchange capacity make it an important material for CO2 adsorption. Moreover, high cation density in the interlayers of montmorillonite also makes it viable for modification to improve its adsorption capabilities furthermore [183,247]. Smectite type minerals can sorb CO2 molecules not only on the surface but also in the interlayer space. However, the content of water molecules and the type of interlaminar cations can play a crucial role in the CO2 sorption capacity [99]. For example, in palygorskite and sepiolite, the dominant mechanism of CO2 sorption is related to their morphological properties and fibrous structure with nanochannels. These channels function as a molecular sieve, where the dimension of cavities is in the appropriate size for retention of CO2 molecules by electrostatic interactions [119,272].
Experiments indicate that illite minerals tend to selectively adsorb CO2 over CH4 when potassium cations are present as exchangeable cations, whereas the impact of cation exchange capacity seems to be more pronounced at low pressure, thus indicating a potential for clay modification [119,266,273,274,275].
Kaolinite has one of the lowest sorption capacities on CO2 capture. CO2 sorption occurs only on the external surface of kaolinite. This sorption happens through the reaction of hydroxide groups within kaolinite and CO2 molecules, which is forming bicarbonate ions [162].
The approach for modification depends essentially on the properties and structure of certain minerals, and uniform strategy here is ambiguous. For example, in regard to amino acid immobilization, clays intercalated with arginine and L-histidine adsorb more CO2 than if they are intercalated with glycine [137]. In addition, amino acid intercalated clay materials are capable of CO2 retention due to the presence of amine groups in the amino acid structure. Unfortunately, the exact mechanisms leading to the retention of CO2 inside the clay mineral structure have not been investigated yet. According to one hypothesis, the adsorption of CO2 can continue through a reaction of a CO2 molecule with an amine group creating carbamic acid group, which can transform to carbamate due to dissociation [137].
Methods for atmospheric CO2 emission reduction in addition to direct adsorption by clay minerals include numerous other approaches (Table 4) with variable scalability and potential costs.
Methods for CO2 emission reduction generally include its fixation, and where possible, it also includes its use, such as in building materials. Carbon pricing is among the primary policies to address climate change and approaches for CO2 emission reduction [277]. Carbon pricing strategies are considered efficient means of reducing carbon emissions [278,279]. However, as of 2019, existing carbon pricing schemes only cover about 20% of global emissions, and more than two-thirds of these have prices below $20 per ton of CO2 equivalent [277]. Thus, an alternative approach (carbon neutrality) that targets fundamental transformation of existing sociotechnical systems is crucial [280].

12. Conclusions

Carbon capture by clay minerals is among the most reasonable options to date as it requires little energy when compared to other techniques and it has comparatively lower development and maintenance costs. Moreover, clay minerals are low-cost, versatile, and abundant in nature, which allow them to be applied on industrial-scale projects, and moreover their applications in most cases are environmentally friendly. The prospects of clay minerals from the Baltic States for industrial-scale carbon capture are thus promising due to the wide availability of clay in the region. The applicability of clay minerals for selective carbon dioxide adsorption can be improved by their modification; suggestions for clay-based carbon dioxide sorbent designs include immobilization of active sorbents onto clay minerals and clay surface modification. In fact, clay minerals are crucial not only in applications for carbon dioxide capture but also in the caprock of geological structures where they contribute towards the immobilization of carbon dioxide through the pore network and thus minimize possible leakage from a potential storage site. Introduction of industrial-scale CO2 capture by clay minerals is a step towards a carbon-neutral future.

Author Contributions

Conceptualization, J.K., M.K. and R.O.-D.; methodology, J.K. and M.K.; investigation, J.K., R.O.-D. and L.A.-B.; resources, J.K., M.K., R.O.-D. and L.A.-B.; data curation, J.K. and M.K.; writing—original draft preparation, J.K., M.K., R.O.-D. and L.A.-B.; writing—review and editing, J.K. and M.K.; supervision, M.K.; project administration, M.K.; funding acquisition, M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the European Regional Development Fund project “Innovation of the waste-to-energy concept for the low carbon economy: development of novel carbon capture technology for thermochemical processing of municipal solid waste (carbon capture and storage from waste—CCSW)”, grant number 1.1.1.1/19/A/013.

Acknowledgments

The authors gratefully acknowledge Juris Burlakovs from Latvian Clay Science Society for effective consulting in clay geology.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Atmospheric Carbon Dioxide Growth Rate. Available online: http://mlg.eng.cam.ac.uk/carl/words/carbon.html (accessed on 24 November 2021).
  2. Global Climate Report—Annual 2020. Available online: https://www.ncdc.noaa.gov/sotc/global/202013 (accessed on 24 November 2021).
  3. Wildfires and Climate Change. Available online: https://www.c2es.org/content/wildfires-and-climate-change/ (accessed on 24 November 2021).
  4. Allen, L.H. Effects of increasing carbon dioxide levels and climate change on plant growth. In Proceedings of a Colloquium “Managing Water Resources in the West under Conditions Climate Uncertainty”, Scottsdale, AZ, USA, 14–16 November 1990; The National Academies Press: Washington, DC, USA, 1991; pp. 101–147. [Google Scholar] [CrossRef]
  5. Taub, D.R. Effects of rising atmospheric concentrations of carbon dioxide on plants. Nat. Educ. Knowl. 2010, 3, 21. [Google Scholar]
  6. Lemordant, L.; Gentine, P.; Swann, A.S.; Cook, B.I.; Scheff, J. Critical impact of vegetation physiology on the continental hydrological cycle in response to increasing CO2. Proc. Natl. Acad. Sci. USA 2018, 115, 4093–4098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. D’Amato, G.; Cecchi, L.; D’Amato, M.; Annesi-Maesano, I. Climate change and respiratory diseases. ERR 2014, 23, 161–169. [Google Scholar] [CrossRef] [PubMed]
  8. Seebacher, F.; Post, E. Climate change impacts on animal migration. Clim. Chang. Responses 2015, 2, 5. [Google Scholar] [CrossRef] [Green Version]
  9. Al-Ghussain, L. Global warming: Review on driving forces and mitigation. Environ. Prog. Sustain. Energy 2019, 38, 13–21. [Google Scholar] [CrossRef] [Green Version]
  10. Trends in Atmospheric Carbon Dioxide. Available online: https://gml.noaa.gov/ccgg/trends/ (accessed on 18 October 2021).
  11. Fawzy, S.; Osman, A.I.; Doran, J.; Rooney, D.W. Strategies for mitigation of climate change: A review. Environ. Chem. Lett. 2020, 18, 2069–2094. [Google Scholar] [CrossRef]
  12. A European Green Deal. Available online: https://ec.europa.eu/info/strategy/priorities-2019-2024/european-green-deal_en (accessed on 1 November 2021).
  13. Outcomes of the Glasgow Climate Change Conference—Advance Unedited Versions (AUVs). Available online: https://unfccc.int/process-and-meetings/conferences/glasgow-climate-change-conference-october-november-2021/outcomes-of-the-glasgow-climate-change-conference (accessed on 17 November 2021).
  14. Alami, A.H.; Hawili, A.A.; Tawalbeh, M.; Hasan, R.; Al Mahmoud, L.; Chibib, S.; Aokal, K.; Mahmood, A.; Rattanapanya, P. Materials and logistics for carbon dioxide capture, storage and utilization. Sci. Total Environ. 2020, 717, 137221. [Google Scholar] [CrossRef]
  15. Boot-Handford, M.E.; Abanades, J.C.; Anthony, E.J.; Blunt, M.J.; Brandani, S.; Mac Dowell, N.; Fernandez, J.R.; Ferrari, M.-C.; Gross, R.; Hallett, J.P.; et al. Carbon capture and storage update. Energy Environ. Sci. 2014, 7, 130–189. [Google Scholar] [CrossRef]
  16. Krumins, J.; Klavins, M.; Delina, A.; Damkevics, R.; Seglins, V. Potential of the middle Cambrian aquifer for carbon dioxide storage in the Baltic States. Energies 2021, 14, 3681. [Google Scholar] [CrossRef]
  17. Pant, D.; Nadda, A.K.; Pant, K.K.; Agarwal, A.K. Advances in Carbon Capture and Utilization. In Advances in Carbon Capture and Utilization; Pant, D., Nadda, A.K., Pant, K.K., Agarwal, A.K., Eds.; Springer: Singapore, 2021; pp. 3–7. [Google Scholar]
  18. Sabri, M.A.; Al Jitan, S.; Bahamon, D.; Vega, L.F.; Palmisano, G. Current and future perspectives on catalytic-based integrated carbon capture and utilization. Sci. Total Environ. 2021, 790, 148081. [Google Scholar] [CrossRef]
  19. Hildenbrand, A.; Schlomer, S.; Krooss, B.M. Gas breakthrough experiments on fine-grained sedimentary rocks. Geofluids 2002, 2, 3–23. [Google Scholar] [CrossRef]
  20. Cole, D.R.; Chialvo, A.A.; Rother, G.; Vlcek, L.; Cummings, P.T. Supercritical fluid behaviour at nanoscale interfaces: Implications for CO2 sequestration in geologic formations. Phil. Mag. 2010, 90, 2339–2363. [Google Scholar] [CrossRef]
  21. Wollenweber, J.; Alles, S.; Busch, A.; Krooss, B.M.; Stanjek, H.; Littke, R. Experimental investigation of the CO2 sealing efficiency of caprocks. Int. J. Greenh. Gas Control 2010, 4, 231–241. [Google Scholar] [CrossRef]
  22. Nesse, W.D. Introduction to Mineralogy, 3rd ed.; Oxford University Press: New York, NY, USA, 2016. [Google Scholar]
  23. Hower, J.; Eslinger, E.V.; Hower, M.E.; Perry, E.A. Mechanism of burial metamorphism of argillaceous sediment: 1. Mineralogical and chemical evidence. GSA Bull. 1976, 87, 725–737. [Google Scholar] [CrossRef]
  24. de Jong, S.M.; Spiers, C.J.; Busch, A. Development of swelling strain in smectite clays through exposure to carbon dioxide. Int. J. Greenh. Gas Control 2014, 24, 149–161. [Google Scholar] [CrossRef]
  25. Dazas, B.; Lanson, B.; Breu, J.; Robert, J.-L.; Pelletier, M.; Ferrage, E. Smectite fluorination and its impact on interlayer water content and structure: A way to fine tune the hydrophilicity of clay surfaces? Microporous Mesoporous Mater. 2013, 181, 233–247. [Google Scholar] [CrossRef]
  26. Franco, F.; Pozo, M.; Cecilia, J.A.; Benitez-Guerrero, M.; Lorente, M. Effectiveness of microwave assisted acid treatment on dioctahedral and trioctahedral smectites. The influence of octahedral composition. Appl. Clay Sci. 2016, 120, 70–80. [Google Scholar] [CrossRef]
  27. Kloproge, J.T. Synthesis of smectites and porous pillared clay catalysts: A review. J. Porous Mater. 1998, 5, 5–41. [Google Scholar] [CrossRef]
  28. Michels, L.; Fossum, J.O.; Rozynek, Z.; Hemmen, H.; Rustenberg, K.; Sobas, P.A.; Kalantzopoulos, G.N.; Knudsen, K.D.; Janek, M.; Plivelic, T.S.; et al. Intercalation and retention of carbon dioxide in a smectite clay promoted by interlayer cations. Sci. Rep. 2015, 5, 8775. [Google Scholar] [CrossRef] [Green Version]
  29. Azzouz, A.; Nousir, S.; Platon, N.; Ghomari, K.; Shiao, T.C.; Hersant, G.; Bergron, J.-Y.; Roy, R. Truly reversible capture of CO2 by montmorillonite intercalated with soya oil-derived polyglycerols. Int. J. Greenh. Gas Control 2013, 17, 140–147. [Google Scholar] [CrossRef]
  30. Druckman, J.N.; McGrath, M.C. The evidence for motivated reasoning in climate change preference formation. Nat. Clim. Chang. 2019, 9, 111–119. [Google Scholar] [CrossRef]
  31. Nielsen, K.S.; van der Linden, S.; Stern, P.C. How behavioural interventions can reduce the climate impact of energy use. Joule 2020, 4, 1613–1616. [Google Scholar] [CrossRef]
  32. van der Linden, S. The social-psychological determinants of climate change risk perceptions, attitudes, and behaviours: A national study. Environ. Educ. Res. 2016, 22, 434–435. [Google Scholar] [CrossRef]
  33. Impact of Covid-19 Crisis on Industrial Production. Available online: https://ec.europa.eu/eurostat/statistics-explained/index.php?title=Impact_of_Covid-19_crisis_on_industrial_production (accessed on 24 November 2021).
  34. Impact of COVID-19 on CO2 Emissions. Available online: https://unfccc.int/sites/default/files/resource/1.GCP_.pdf (accessed on 18 October 2021).
  35. Largest Producers of Fossil Fuel CO2 Emissions Worldwide in 2019, by Share of Emissions. Available online: https://www.statista.com/statistics/271748/the-largest-emitters-of-co2-in-the-world/ (accessed on 27 October 2021).
  36. Keeling Curve. Available online: https://www.climatecentral.org/gallery/graphics/keeling_curve (accessed on 18 October 2021).
  37. Global CO2 Emissions Rebound by Nearly 5% in 2021, Approaching the 2018–2019 Peak. Available online: https://www.iea.org/reports/global-energy-review-2021/co2-emissions (accessed on 3 November 2021).
  38. Statistical Review of World Energy 2021. Available online: https://www.bp.com/content/dam/bp/business-sites/en/global/corporate/pdfs/energy-economics/statistical-review/bp-stats-review-2021-full-report.pdf (accessed on 29 November 2021).
  39. Climate Change Mitigation Policies in Energy Sector of Baltic States. Available online: https://www.worldenergy.org/assets/downloads/PUB_Energy_and_Climate_Change_Annex_Baltic_states_2007_WEC.pdf (accessed on 15 November 2021).
  40. Greenhouse Gas Emissions by Source Sector. Available online: http://appsso.eurostat.ec.europa.eu/nui/submitViewTableAction.do (accessed on 17 November 2021).
  41. Font-Palma, C.; Cann, D.; Udemu, C. Review of cryogenic carbon capture innovations and their potential applications. C 2021, 7, 58. [Google Scholar] [CrossRef]
  42. Lin, C.-C.; Liu, W.-T.; Tan, C.-S. Removal of carbon dioxide by absorption in a rotating packed bed. Ind. Eng. Chem. Res. 2003, 42, 2381–2386. [Google Scholar] [CrossRef]
  43. Dave, A.; Pathak, B.; Dave, M.; Rezvani, S.; Huang, Y.; Hewitt, N. Process design of CO2 desorption from physical solvent di-methyl-ether of poly-ethylene-glycol. Mater. Sci. Energy Technol. 2020, 3, 209–2017. [Google Scholar] [CrossRef]
  44. Yang, S.; Qian, Y.; Yang, S. Development of a full CO2 capture process based on the rectisol wash technology. Ind. Eng. Chem. Res. 2016, 55, 6186–6193. [Google Scholar] [CrossRef]
  45. Scherffius, J.R.; Reddy, S.; Klumpyan, J.P.; Armpriester, A. Large-scale CO2 capture demonstration plant using fluor’s econamine FG Plus technology at NRG’s WA parish electric generating station. Energy Procedia 2013, 37, 6553–6561. [Google Scholar] [CrossRef] [Green Version]
  46. Hochgesand, G. Rectisol and purisol. Ind. Eng. Chem. 1970, 62, 37–43. [Google Scholar] [CrossRef]
  47. Girimonte, R.; Testa, F.; Gallo, F.; Buscieti, R.; Leone, G.; Fromisani, B. Adsorption of CO2 on amine-modified silica particles in a confined-fluidized bed. Processes 2020, 8, 1531. [Google Scholar] [CrossRef]
  48. Auta, M.; Amat Darbis, N.D.; Mohd Din, A.T.; Harmeed, B.H. Fixed-bed column adsorption of carbon dioxide by sodium hydroxide modified activated alumina. Chem. Eng. J. 2013, 233, 80–87. [Google Scholar] [CrossRef]
  49. Casas, N.; Schell, J.; Pini, R.; Mazzotti, M. Fixed bed adsorption of CO2/H2 mixtures on activated carbon: Experiments and modelling. Adsorption 2012, 18, 143–161. [Google Scholar] [CrossRef] [Green Version]
  50. Hauchhum, L.; Mahanta, P. Carbon dioxide adsorption on zeolites and activated carbon by pressure swing adsorption in a fixed bed. Int. J. Energy Environ. Eng. 2014, 5, 349–356. [Google Scholar] [CrossRef] [Green Version]
  51. Bastin, L.; Barcia, P.S.; Hurtado, E.J.; Silva, J.A.C.; Rodrigues, A.E.; Chen, B. A microporous metal−organic framework for separation of CO2/N2 and CO2/CH4 by fixed-bed adsorption. J. Phys. Chem. C 2008, 112, 1575–1581. [Google Scholar] [CrossRef]
  52. Moreira, R.F.P.M.; Soares, J.L.; Casarin, G.L.; Rodrigues, A.E. Adsorption of CO2 on hydrotalcite-like compounds in a fixed bed. Sep. Sci. Technol. 2006, 41, 341–357. [Google Scholar] [CrossRef]
  53. Guo, Y.; Luo, L.; Zheng, Y.; Zhu, T. Optimization of CO2 adsorption on solid-supported Amines and thermal regeneration mode comparison. ACS Omega 2020, 5, 9641–9648. [Google Scholar] [CrossRef] [Green Version]
  54. Yoro, K.O.; Singo, M.; Mulopo, J.L.; Daramola, M.O. Modelling and experimental study of the CO2 adsorption behaviour of polyaspartamide as an adsorbent during post-combustion CO2 capture. Energy Procedia 2017, 114, 1643–1664. [Google Scholar] [CrossRef]
  55. Siqueira, R.M.; Freitas, G.R.; Peixoto, H.R.; do Nascimento, J.F.; Musse, A.P.S.; Torres, A.E.B.; Azevedo, D.C.S.; Bastos-Neto, M. Carbon dioxide capture by pressure swing adsorption. Energy Procedia 2017, 114, 2182–2192. [Google Scholar] [CrossRef]
  56. Ntiamoah, A.; Ling, J.; Xiao, P.; Webley, P.A.; Zhai, Y. CO2 capture by temperature swing adsorption: Use of hot CO2-rich gas for regeneration. Ind. Eng. Chem. Res. 2016, 55, 703–713. [Google Scholar] [CrossRef]
  57. Elliott, J.E.; Copeland, R.J.; Leta, D.P.; McCall, P.P.; Bai, C.; DeRites, B.A. Carbon Dioxide Separation Using Adsorption with Steam Regeneration. US Patent 9504955, 29 November 2016. [Google Scholar]
  58. Dean, C.C.; Blamey, J.; Florin, N.H.; Al-Jeboori, M.J.; Fennell, P.S. The calcium looping cycle for CO2 capture from power generation, cement manufacture and hydrogen production. Chem. Eng. Res. Des. 2011, 89, 836–855. [Google Scholar] [CrossRef] [Green Version]
  59. Xu, J.; He, T.; Lin, W. Experimental and theoretical study of CO2 solubility in liquid CH4/H2 mixtures at cryogenic temperatures. J. Chem. Eng. Data 2021, 66, 2844–2855. [Google Scholar] [CrossRef]
  60. Chang, H.-M.; Chung, M.J.; Park, S.B. Cryogenic heat-exchanger design for freeze-out removal of carbon dioxide from landfill gas. J. Therm. Sci. Technol. 2009, 4, 362–371. [Google Scholar] [CrossRef] [Green Version]
  61. Babar, M.; Bustam, M.A.; Maulud, A.S.; Ali, A.; Mukhtar, A.; Ullah, S. Enhanced cryogenic packed bed with optimal CO2 removal from natural gas; a joint computational and experimental approach. Cryogenics 2020, 105, 103010. [Google Scholar] [CrossRef]
  62. Zhimin, H.; Zhigang, T.; Ataeivarjovi, E.; Dong, G.; Zhijun, Z.; Hongwei, L. Study on polydimethylsiloxane desorption membrane of CO2-dimethyl carbonate system. Energy Procedia 2017, 118, 210–215. [Google Scholar] [CrossRef]
  63. deMontigny, D.; Tontiwachwuthikul, P.; Chakma, A. Using polypropylene and polytetrafluoroethylene membranes in a membrane contactor for CO2 absorption. J. Membr. Sci. 2006, 277, 99–107. [Google Scholar] [CrossRef]
  64. Drioli, E.; Brunetti, A.; Barbieri, G. Ceramic membranes in carbon dioxide capture: Applications and potentialities. Adv. Sci. Technol. 2010, 72, 105–118. [Google Scholar] [CrossRef]
  65. Widayat, W.; Suzery, M.; Satriadi, H.; Wahyudi; Philia, J. Selection and cultivation of microalgae for CO2 biofixation. IOP Conf. Ser. Mater. Sci. Eng. 2021, 1053, 012132. [Google Scholar] [CrossRef]
  66. Xu, X.; Kentish, S.E.; Martin, G.J.O. Direct air capture of CO2 by microalgae with buoyant beads encapsulating carbonic anhydrase. ACS Sustain. Chem. Eng. 2021, 9, 9698–9706. [Google Scholar] [CrossRef]
  67. Pierre, A.C. Enzymatic carbon dioxide capture. Int. Sch. Res. Notices 2012, 2012, 753687. [Google Scholar] [CrossRef]
  68. Choi, S.; Dresse, J.H.; Jones, C.W. Adsorbent materials for carbon dioxide capture from large anthropogenic point sources. ChemSusChem 2009, 2, 796–854. [Google Scholar] [CrossRef]
  69. Pellegrini, L.A.; De Guido, G.; Ingrosso, S. Thermodynamic Framework for Cryogenic Carbon Capture. Comput. Aided Chem. Eng. 2020, 48, 475–480. [Google Scholar] [CrossRef]
  70. Shah, S.; Shah, M.; Shah, A.; Shah, M. Evolution in the membrane-based materials and comprehensive review on carbon capture and storage in industries. Emerg. Mater. 2020, 3, 33–44. [Google Scholar] [CrossRef]
  71. Zhang, X.; Song, Z.; Gani, R.; Zhou, T. Comparative economic analysis of physical, chemical, and hybrid absorption processes for carbon capture. Ind. Eng. Chem. Res. 2020, 59, 2005–2012. [Google Scholar] [CrossRef]
  72. Ghanbari, T.; Abnisa, F.; Daud, W.M.A.W. A review on production of metal organic frameworks (MOF) for CO2 adsorption. Sci. Total Environ. 2020, 707, 135090. [Google Scholar] [CrossRef] [PubMed]
  73. Sharma, H.; Dhir, A. Capture of carbon dioxide using solid carbonaceous and non-carbonaceous adsorbents: A review. Environ. Chem. Lett. 2021, 19, 851–873. [Google Scholar] [CrossRef]
  74. Vilarrasa-Garcia, E.; Cecilia, J.A.; Azevedo, D.C.S.; Cavalcante, C.L.; Rodriguez-Castellon, E. Evaluation of porous clay heterostructures modified with amine species as adsorbent for CO2 capture. Microporous Mesoporous Mater. 2017, 249, 25–33. [Google Scholar] [CrossRef]
  75. Park, S.; Wijaya, D.T.; Na, J.; Lee, C.W. Towards the large-scale electrochemical reduction of carbon dioxide. Catalysts 2021, 11, 253. [Google Scholar] [CrossRef]
  76. Pennline, H.W.; Granite, E.J.; Luebke, D.R.; Kitchin, J.R.; Landon, J.; Weiland, L.M. Separation of CO2 from flue gas using electrochemical cells. Fuel 2010, 89, 1307–1314. [Google Scholar] [CrossRef]
  77. Rinker, E.B.; Ashour, S.S.; Sandall, O.C. Absorption of carbon dioxide into aqueous blends of diethanolamine and methyldiethanolamine. Ind. Eng. Chem. Res. 2000, 39, 4346–4356. [Google Scholar] [CrossRef]
  78. Siegelman, R.L.; Kim, E.J.; Long, J.R. Porous materials for carbon dioxide separations. Nat. Mater. 2021, 20, 1060–1072. [Google Scholar] [CrossRef]
  79. Osman, A.I.; Hefny, M.; Abdel Maksoud, M.I.A.; Elgarahy, A.M.; Rooney, D.W. Recent advances in carbon capture storage and utilisation technologies: A review. Environ. Chem. Lett. 2021, 19, 797–849. [Google Scholar] [CrossRef]
  80. Siriwardane, R.V.; Shen, M.-S.; Fisher, E.P.; Poston, J.A. Adsorption of CO2 on molecular sieves and activated carbon. Energy Fuels 2001, 15, 279–284. [Google Scholar] [CrossRef]
  81. Soares, J.L.; Moreira, R.F.P.M.; Jose, H.J.; Grande, C.A.; Rodrigues, A.E. Hydrotalcite materials for carbon dioxide adsorption at high temperatures: Characterization and diffusivity measurements. Sep. Sci. Technol. 2004, 39, 1989–2010. [Google Scholar] [CrossRef]
  82. Millward, A.R.; Yaghi, O.M. Metal-organic frameworks with exceptionally high capacity for storage of carbon dioxide at room temperature. J. Am. Chem. Soc. 2005, 127, 17998–17999. [Google Scholar] [CrossRef]
  83. Walton, K.S.; Abney, M.B.; LeVan, M.D. CO2 adsorption in Y and X zeolites modified by alkali metal cation exchange. Microporous Mesoporous Mater. 2006, 91, 78–84. [Google Scholar] [CrossRef]
  84. Bae, Y.-S.; Mulfort, K.L.; Frost, H.; Ryan, P.; Punnathanam, S.; Broadbelt, L.J.; Hupp, J.T.; Snurr, R.Q. Separation of CO2 from CH4 using mixed-ligand metal-organic frameworks. Langmuir 2008, 24, 8592–8598. [Google Scholar] [CrossRef]
  85. Martavaltzi, C.S.; Lemonidou, A.A. Development of new CaO based sorbent materials for CO2 removal at high temperature. Microporous Mesoporous Mater. 2008, 110, 119–127. [Google Scholar] [CrossRef]
  86. Shao, W.; Zhang, L.; Li, L. Adsorption of CO2 and N2 on synthesized NaY zeolite at high temperatures. Adsorption 2009, 15, 497. [Google Scholar] [CrossRef]
  87. Huang, C.-M.; Hsu, H.-W.; Liu, W.-H.; Cheng, J.-Y.; Chen, W.-C.; Wen, T.-W.; Chen, W. Development of post-combustion CO2 capture with CaO/CaCO3 looping in a bench scale plant. Energy Procedia 2011, 4, 1268–1275. [Google Scholar] [CrossRef] [Green Version]
  88. Bezzera, D.P.; da Silva, F.W.M.; de Moura, P.A.S.; Sousa, A.G.S.; Vieira, R.S.; Rodriguez-Castellon, E.; Azevedo, D.C.S. CO2 adsorption in amine-grafted zeolite 13X. Appl. Surf. Sci. 2014, 314, 314–321. [Google Scholar] [CrossRef] [Green Version]
  89. Loganathan, S.; Tikmani, M.; Edubilli, S.; Mishra, A.; Ghoshal, A.K. CO2 adsorption kinetics on mesoporous silica under wide range of pressure and temperature. Chem. Eng. J. 2014, 256, 1–8. [Google Scholar] [CrossRef]
  90. Pera-Titus, M. Porous Inorganic Membranes for CO2 Capture: Present and Prospects. Chem. Rev. 2014, 114, 1413–1492. [Google Scholar] [CrossRef] [PubMed]
  91. Bhatta, L.K.G.; Subramanyam, S.; Chengala, M.D.; Olivera, S.; Venkatesh, K. Progress in hydrotalcite like compounds and metal-based oxides for CO2 capture: A review. J. Clean. Prod. 2015, 103, 171–196. [Google Scholar] [CrossRef]
  92. Chang, S.-C.; Chien, S.-Y.; Chen, C.-L.; Chen, C.-K. Analysing adsorption characteristics of CO2, N2 and H2O in MCM-41 silica by molecular simulation. Appl. Surf. Sci. 2015, 331, 225–233. [Google Scholar] [CrossRef]
  93. Montagnaro, F.; Silvestre-Albero, A.; Silvestre-Albero, J.; Rodriguez-Reinoso, F.; Erto, A.; Lancia, A.; Balsamo, M. Post-combustion CO2 adsorption on activated carbons with different textural properties. Microporous Mesoporous Mater. 2015, 209, 157–164. [Google Scholar] [CrossRef]
  94. Vilarrasa-Garcia, E.; Cecilia, J.A.; Santos, S.M.L.; Cavalcante, C.L.; Jimenez-Jimenez, J.; Azevedo, D.C.S.; Rodriguez-Castellon, E. CO2 adsorption on APTES functionalized mesocellular foams obtained from mesoporous silicas. Microporous Mesoporous Mater. 2014, 187, 125–134. [Google Scholar] [CrossRef]
  95. Vilarrasa-Garcia, E.; Moya, E.M.O.; Cecilia, J.A.; Cavalcante, C.L.; Jimenez-Jimenez, J.; Azevedo, D.C.S.; Rodriguez-Castellon, E. CO2 adsorption on amine modified mesoporous silicas: Effect of the progressive disorder of the honeycomb arrangement. Microporous Mesoporous Mater. 2015, 209, 172–183. [Google Scholar] [CrossRef]
  96. Wang, Q.; Huang, L. Molecular insight into competitive adsorption of methane and carbon dioxide in montmorillonite: Effect of clay structure and water content. Fuel 2019, 239, 32–43. [Google Scholar] [CrossRef]
  97. Stanly, S.; Jelmy, E.J.; Nair, C.P.R.; John, H. Carbon dioxide adsorption studies on modified montmorillonite clay/reduced graphene oxide hybrids at low pressure. J. Environ. Chem. Eng. 2019, 7, 103344. [Google Scholar] [CrossRef]
  98. Samanta, A.; Zhao, A.; Shimizu, G.K.H.; Sarkar, P.; Gupta, R. Post-Combustion CO2 capture using solid sorbents: A Review. Ind. Eng. Chem. Res. 2012, 51, 1438–1463. [Google Scholar] [CrossRef]
  99. Chouikhi, N.; Cecilia, J.A.; Vilarrasa-Garcia, E.; Besghaier, S.; Chlendi, M.; Duro, F.I.F.; Castellon, R.R.; Bagane, M. CO2 adsorption of materials synthesized from clay minerals: A review. Minerals 2019, 9, 514. [Google Scholar] [CrossRef] [Green Version]
  100. Luedtke, Z.; Aro, M.; Sun, Z.; Toan, S. Carbon Capture Materials and Technologies: A Review. Cur. Res. Mater. Chem. 2021, 3, 108. [Google Scholar] [CrossRef]
  101. Aniruddha, R.; Sreedhar, I.; Reddy, B.M. MOFs in carbon capture: Past, present and future. J. CO2 Util. 2020, 42, 101297. [Google Scholar] [CrossRef]
  102. Mutch, G.A.; Shulda, S.; McCue, A.J.; Menart, M.J.; Ciobanu, C.V.; Ngo, C.; Anderson, J.A.; Richards, R.M.; Vega-Maza, D. Carbon capture by metal oxides: Unleashing the potential of the (111) facet. J. Am. Chem. Soc. 2018, 140, 4736–4742. [Google Scholar] [CrossRef] [Green Version]
  103. Zhang, X.; Zhang, X.; Dong, H.; Zhao, Z.; Zhang, S.; Huang, Y. Carbon capture with ionic liquids: Overview and progress. Energy Environ. Sci. 2012, 5, 6668–6681. [Google Scholar] [CrossRef]
  104. Sumida, K.; Rogow, D.L.; Mason, J.A.; McDonald, T.M.; Bloch, E.D.; Herm, Z.R.; Bae, T.H.; Long, J.R. Carbon dioxide capture in metal-organic frameworks. Chem. Rev. 2012, 112, 724–781. [Google Scholar] [CrossRef]
  105. Haque, E.; Islam, M.; Pourazadi, E.; Sarkar, S.; Harris, A.T.; Minett, A.I.; Yanmaz, E.; Alshehri, S.M.; Ide, Y.; Wu, K.C.-W.; et al. Boron functionalized graphene oxide-organic frameworks for highly efficient CO2 capture. Chem. Asian J. 2017, 12, 283–288. [Google Scholar] [CrossRef]
  106. Elhenawy, S.E.M.; Khraisheh, M.; AlMomani, F.; Walker, G. Metal-organic frameworks as a platform for CO2 capture and chemical processes: Adsorption, membrane separation, catalytic-conversion, and electrochemical reduction of CO2. Catalysts 2020, 10, 1293. [Google Scholar] [CrossRef]
  107. Wang, W.; Xiao, J.; Wei, X.; Ding, J.; Wang, X.; Song, C. Development of a new clay supported polyethyleneimine composite for CO2 capture. Appl. Energy 2014, 113, 334–341. [Google Scholar] [CrossRef]
  108. Yang, L.; Zeng, X.; Wang, W.; Cao, D. Recent progress in MOF-derived, heteroatom-doped porous carbons as highly efficient electrocatalysts for oxygen reduction reaction in fuel cells. Adv. Funct. Mat. 2018, 28, 1704537. [Google Scholar] [CrossRef]
  109. Ahmed, I.; Jhung, S.H. Composites of metal–organic frameworks: Preparation and application in adsorption. Mater. Today 2014, 17, 136–146. [Google Scholar] [CrossRef]
  110. Inoue, T.; Fujishima, A.; Konishi, S.; Honda, K. Photo electrocatalytic reduction of carbon dioxide in aqueous suspensions of semiconductor powders. Nature 1979, 277, 637–638. [Google Scholar] [CrossRef]
  111. Wang, S.; Yao, W.; Lin, J.; Ding, Z.; Wang, X. Cobalt imidazolate metal–organic frameworks photosplit CO2 under mild reaction conditions. Angew. Chem. Int. Ed. 2013, 23, 1034–1038. [Google Scholar] [CrossRef]
  112. White, J.L.; Baruch, M.F.; Pander, J.E.; Hu, Y.; Fortmeyer, I.C.; Park, J.E.; Zhang, T.; Liao, K.; Gu, J.; Yan, Y.; et al. Light-driven heterogeneous reduction of carbon dioxide: Photocatalysts and photoelectrodes. Chem. Rev. 2015, 115, 12888–12935. [Google Scholar] [CrossRef]
  113. Wang, S.; Wang, X. Imidazolium ionic liquids, imidazolylidene heterocyclic carbenes, and zeolitic imidazolate frameworks for CO2 capture and photochemical reduction. Angew. Chem. Int. Ed. 2016, 55, 2308–2320. [Google Scholar] [CrossRef]
  114. Lanni, L.M.; Tilford, R.W.; Bharathy, M.; Lavigne, J.J. Enhanced hydrolytic stability of self-assembling alkylated two-dimensional covalent organic frameworks. J. Am. Chem. Soc. 2011, 133, 13975–13983. [Google Scholar] [CrossRef]
  115. Uribe-Romo, F.J.; Doonan, C.J.; Furukawa, H.; Oisaki, K.; Yaghi, O.M. Crystalline covalent organic frameworks with hydrazone linkages. J. Am. Chem. Soc. 2011, 133, 11478–11481. [Google Scholar] [CrossRef]
  116. Li, H.; Li, H.; Dai, Q.; Li, H.; Brédas, J.-L. Hydrolytic stability of boronate ester-linked covalent organic frameworks. Adv. Theory Sim. 2018, 1, 1700015. [Google Scholar] [CrossRef]
  117. Zhang, B.; Wei, M.; Mao, H.; Pei, X.; Alshmimri, S.A.; Reimer, J.A.; Yaghi, O.M. Crystalline dioxin-linked covalent organic frameworks from irreversible reactions. J. Am. Chem. Soc. 2018, 140, 12715–12719. [Google Scholar] [CrossRef] [Green Version]
  118. Ramis, G.; Busca, G.; Lorenzelli, V. Low-temperature CO2 adsorption on metal oxides: Spectroscopic characterization of some weakly adsorbed species. Mater. Chem. Phys. 1991, 29, 425–435. [Google Scholar] [CrossRef]
  119. Cecilia, J.A.; Vilarrasa-García, E.; Cavalcante, C.L.; Azevedo, D.C.S.; Franco, F.; Rodríguez-Castellón, E. Evaluation of two fibrous clay minerals (sepiolite and palygorskite) for CO2 Capture. J. Environ. Chem. Eng. 2018, 6, 4573–4587. [Google Scholar] [CrossRef]
  120. Sanz-Pérez, E.S.; Arencibia, A.; Calleja, G.; San, R. Tuning the textural properties of HMS mesoporous silica. Functionalization towards CO2 adsorption. Microporous Mesoporous Mater. 2018, 260, 235–244. [Google Scholar] [CrossRef]
  121. Gomez-Pozuelo, G.; Sanz-Perez, E.S.; Arencibia, A.; Pizarro, P.; Sanz, R.; Serrano, D.P. CO2 adsorption on amine-functionalized clays. Microporous Mesoporous Mater. 2019, 282, 38–47. [Google Scholar] [CrossRef]
  122. Hiyoshi, N.; Yogo, K.; Yashima, T. Adsorption characteristics of carbon dioxide on organically functionalized SBA-15. Microporous Mesoporous Mater. 2005, 84, 357–365. [Google Scholar] [CrossRef]
  123. Vilarrasa-Garcia, E.; Cecilia, J.A.; Moya, E.M.O.; Cavalcante, C.L.; Azevedo, D.C.S.; Rodriguez-Castellon, E. “Low cost” pore expanded SBA-15 functionalized with amine groups applied to CO2 adsorption. Materials 2015, 8, 2495–2513. [Google Scholar] [CrossRef] [Green Version]
  124. Cecilia, J.A.; Vilarrasa-García, E.; García-Sancho, C.; Saboya, R.M.A.; Azevedo, D.C.S.; Cavalcante, C.L.; Rodríuez-Castellón, E. Functionalization of hollow silica microspheres by impregnation or grafted of amine groups for the CO2 capture. Int. J. Greenh. Gas Control 2016, 52, 344–356. [Google Scholar] [CrossRef]
  125. Yan, X.; Zhang, L.; Zhang, Y.; Qiao, K.; Yan, Z.; Komarnemi, S. Amine-modified mesocellular silica foams for CO2 capture. Chem. Eng. J. 2011, 168, 918–924. [Google Scholar] [CrossRef]
  126. Li, G.; Xiao, P.; Webley, P.; Zhang, J.; Singh, R.; Marshall, M. Capture of CO2 from high humidity flue gas by vacuum swing adsorption with zeolite 13X. Adsorption 2008, 14, 415–422. [Google Scholar] [CrossRef]
  127. Merel, J.; Clausse, M.; Meurier, F. Experimental investigation on CO2 post-combustion capture by indirect thermal swing adsorption using 13X and 5A zeolites. Ind. Eng. Chem. Res. 2008, 47, 209–215. [Google Scholar] [CrossRef]
  128. Plaza, M.G.; Pevida, C.; Arenillas, A.; Rubiera, F.; Pis, J.J. CO2 capture by adsorption with nitrogen enriched carbons. Fuel 2007, 86, 2204–2212. [Google Scholar] [CrossRef] [Green Version]
  129. Pevida, C.; Plaza, M.G.; Arias, B.; Fermoso, J.; Rubiera, F.; Pis, J.J. Surface modification of activated carbons for CO2 capture. Appl. Surf. Sci. 2008, 254, 7165–7172. [Google Scholar] [CrossRef] [Green Version]
  130. Peng, J.; Sun, H.; Wang, J.; Qiu, F.; Zhang, P.; Ning, W.; Zhang, D.; Li, W.; Wei, C.; Miao, S. Highly stable and recyclable sequestration of CO2 using supported melamine on layered-chain clay mineral. Appl. Mater. Interfaces 2021, 13, 10933–10941. [Google Scholar] [CrossRef] [PubMed]
  131. Cavalcanti, L.P.; Kalantzopoulos, N.G.; Eckert, J.; Knudsen, K.D.; Fossum, J.O. A nano-silicate material with exceptional capacity for CO2 capture and storage at room temperature. Sci. Rep. 2018, 8, 11827. [Google Scholar] [CrossRef] [PubMed]
  132. Wang, W.; Wang, X.; Song, C.; Wei, X.; Ding, J.; Xiao, J. Sulfuric acid modified bentonite as the support of tetraethylenepentamine for CO2 capture. Energy Fuels 2013, 27, 1538–1546. [Google Scholar] [CrossRef]
  133. Teng, Y.; Liu, Z.; Xu, G.; Zhang, K. Desorption kinetics and mechanisms of CO2 on amine-based mesoporous silica materials. Energies 2017, 10, 115. [Google Scholar] [CrossRef]
  134. Shukla, S.K.; Khokarale, S.G.; Bui, T.Q.; Mikkola, J.-P.T. Ionic liquids: Potential materials for carbon dioxide capture and utilization. Front. Mater. 2019, 6, 42. [Google Scholar] [CrossRef] [Green Version]
  135. Steinruck, H.-P.; Wasserscheid, P. Ionic liquids in catalysis. Catal. Lett. 2015, 145, 380–397. [Google Scholar] [CrossRef]
  136. DePaolo, D.J.; Cole, D.R. Geochemistry of geologic carbon sequestration: An overview. Rev. Mineral. Geochem. 2013, 77, 1–14. [Google Scholar] [CrossRef]
  137. Juzkow, J. Clay Materials Modified with Amino Acids for Purification Processes of Biogas and Natural Gas. Master’s Thesis, Technical University of Lisbon, Lisbon, Portugal, 2016. [Google Scholar]
  138. Vaccari, A. Clays and catalysis: A promising future. Appl. Clay Sci. 1999, 14, 161–198. [Google Scholar] [CrossRef]
  139. An, N.; Zhou, C.H.; Zhuang, X.Y.; Tong, D.S.; Yu, W.H. Immobilization of enzymes on clay minerals for biocatalysts and biosensors. Appl. Clay Sci. 2015, 114, 283–296. [Google Scholar] [CrossRef]
  140. Guggenheim, S.; Martin, R.T. Definition of clay and clay mineral: Joint report of the AIPEA nomenclature and CMS nomenclature committees. Clays Clay Miner. 1995, 43, 255–256. [Google Scholar] [CrossRef]
  141. Schulze, D.G. Clay minerals. In Encyclopaedia of Soils in the Environment; Hillel, D., Ed.; Elsevier: Boston, MA, USA, 2005; pp. 246–254. [Google Scholar]
  142. Theng, B.K.G. The clay minerals. Dev. Clay Sci. 2012, 4, 3–45. [Google Scholar] [CrossRef]
  143. Ismadji, S.; Soetaredjo, F.E.; Ayucitra, A. Natural clay minerals as environmental cleaning agents. In Clay Materials for Environmental Remediation; Ismadji, S., Soetaredjo, F.E., Ayucitra, A., Eds.; Springer: Copenhagen, Denmark, 2015; pp. 5–37. [Google Scholar] [CrossRef]
  144. Christidis, G.E. Industrial clays. In European Mineralogical Union Notes in Mineralogy. Advances in the Characterization of Industrial Clays; Christidis, G.E., Ed.; Mineralogical Society of Great Britain and Ireland: London, UK, 2010; Volume 9, pp. 341–414. [Google Scholar] [CrossRef]
  145. Schoonheydt, R.A.; Johnston, C.T.; Bergaya, F. Clay minerals and their surfaces. Dev. Clay Sci. 2018, 9, 1–21. [Google Scholar] [CrossRef]
  146. Clay Mineral. Available online: https://www.britannica.com/science/clay-mineral (accessed on 7 October 2021).
  147. Zhou, C.H.; Keeling, J. Fundamental and applied research on clay minerals: From climate and environment to nanotechnology. App. Clay Sci. 2013, 74, 3–9. [Google Scholar] [CrossRef]
  148. Brigatti, M.F.; Galan, E.; Theng, B.K.H. Structures and mineralogy of clay minerals. Dev. Clay Sci. 2006, 1, 19–86. [Google Scholar] [CrossRef]
  149. Brigatti, M.F.; Malferrari, D.; Laurora, A.; Elmi, C.; Mottana, A. Structure and mineralogy of layer silicates: Recent perspectives and new trends. Layered mineral structures and their application in advanced technologies. EMU Notes Mineral. 2011, 11, 1–71. [Google Scholar]
  150. Martin, R.T.; Bailey, S.W.; Eberl, D.D.; Fanning, D.S.; Guggenheim, S.; Kodama, H.; Pevear, D.R.; Srodon, J.; Wicks, F.J. Report of the Clay Minerals Society Nomenclature Committee: Revised classification of clay materials. Clays Clay Miner. 1991, 39, 333–335. [Google Scholar] [CrossRef]
  151. Guggenheim, S.; Adams, J.M.; Bain, D.C.; Bergaya, F.; Brigatti, M.F.; Drits, V.A.; Formoso, M.L.L.; Galán, E.; Kogure, T.; Stanjek, H. Summary of recommendations of Nomenclature Committees relevant to clay mineralogy: Report of the association Internationale pour l’Etude des Argiles (AIPEA) Nomenclature Committee for 2006. Clays Clay Miner. 2006, 54, 761–772. [Google Scholar] [CrossRef]
  152. Schroeder, P. Clays in the Critical Zone; Cambridge University Press: Cambridge, UK, 2018; pp. 1–246. [Google Scholar] [CrossRef]
  153. Bergaya, F.; Lagaly, G. General introduction: Clays, clay minerals, and clay science. Dev. Clay. Sci. 2006, 1, 1–18. [Google Scholar] [CrossRef]
  154. Boulet, P.; Greenwell, H.C.; Stackhouse, S.; Coveney, P.V. Recent advances in understanding the structure and reactivity of clays using electronic structure calculations. J. Mol. Struc. Theochem 2006, 762, 33–48. [Google Scholar] [CrossRef]
  155. da Silva, G.J.; Fossum, J.O.; DiMasi, E.; Maloy, K.J.; Lutnaes, S.B. Synchrotron x-ray scattering studies of water intercalation in a layered synthetic silicate. Phys. Rev. E 2002, 66, 011303. [Google Scholar] [CrossRef] [PubMed]
  156. Bordallo, H.N.; Aldridge, L.P.; Churchman, G.J.; Gates, W.P.; Telling, M.T.F.; Kiefer, K.; Fouquet, P.; Seydel, T.; Kimber, S.A.J. Quasi-elastic neutron scattering studies on clay interlayer-space highlighting the effect of the cation in confined water dynamics. J. Phys. Chem. C. 2008, 112, 13982–13991. [Google Scholar] [CrossRef]
  157. Hansen, E.L.; Hemmen, H.; Fonseca, D.M.; Coutant, C.; Knudsen, K.D.; Plivelic, T.S.; Bonn, D.; Fossum, J.O. Swelling transition of a clay induced by heating. Sci. Rep. 2012, 2, 618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Murray, H.H. Applied Clay Mineralogy: Occurrences, Processing and Application of Kaolins, Bentonites, Palygorskite-Sepiolite, and Common Clays, 1st ed.; Elsevier: Amsterdam, The Netherlands, 2006; pp. 1–180. [Google Scholar]
  159. Yariv, S.; Cross, H. Organo-Clay Complexes and Interactions; Marcel Dekker: New York, NY, USA, 2002; pp. 1–688. [Google Scholar]
  160. Stinkule, A. Clays in the Subterranean of Latvia; RTU Publishing House: Riga, Latvia, 2014; pp. 1–121. [Google Scholar]
  161. Lan, Y.; Liu, Y.; Li, J.; Chen, D.; He, G.; Parkin, I.P. Natural clay-based materials for energy storage and conversion applications. Adv. Sci. 2021, 8, 2004036. [Google Scholar] [CrossRef]
  162. Chen, Y.; Lu, D. CO2 capture by kaolinite and its adsorption mechanism. Appl. Clay Sci. 2015, 104, 221–228. [Google Scholar] [CrossRef]
  163. Birbara, P.J.; Nalette, T.A. Regenerable Supported Amine-Polyol Sorbent. US Patent 5,376,614, 27 December 1994. [Google Scholar]
  164. Yong, Z.; Rodrigues, A.E. Hydrotalcite-like compounds as adsorbents for carbon dioxide. Energy Convers. Manag. 2002, 43, 1865–1876. [Google Scholar] [CrossRef]
  165. Gray, M.L.; Soong, Y.; Champagne, K.J.; Baltrus, J.; Stevens, R.W.; Toochinda, P.; Chuang, S.S.C. CO2 capture by amine-enriched fly ash carbon sorbents. Sep. Purif. Technol. 2004, 35, 31–36. [Google Scholar] [CrossRef]
  166. Santos-Costa, A.; Imae, T.; Takagi, K.; Kikuta, K. Intercalation of dendrimers in the interlayer of hydrotalcite clay sheets. In Surface and Colloid Science; Springer: Berlin/Heidelberg, Germany, 2004; Volume 128, pp. 113–119. [Google Scholar] [CrossRef]
  167. Sirwardane, R.V. Solid sorbents for removal of carbon dioxide from gas streams at low temperatures. US Patent 6908497, 21 June 2005. [Google Scholar]
  168. Chaffee, A.L.; Knowles, G.P.; Liang, Z.; Zhang, J.; Xiao, P.; Webley, P.A. CO2 capture by adsorption: Materials and process development. Int. J. Greenh. Gas Control 2007, 1, 11–18. [Google Scholar] [CrossRef]
  169. Thomas, J.; Bohor, B.F. Surface area of vermiculite with nitrogen and carbon dioxide as adsorbates. Clays Clay Miner. 1969, 17, 205–209. [Google Scholar] [CrossRef]
  170. Aylmore, L.A.G. Gas sorption in clay mineral systems. Clays Clay Miner. 1974, 22, 175–183. [Google Scholar] [CrossRef]
  171. Azzouz, A. Physicochimie des Tamis Moléculaires; Office des Publications Universitaires: Algiers, France, 1994. [Google Scholar]
  172. Rodlert, M.; Plummer, C.J.G.; Leterrier, Y.; Manson, J.-A.E.; Grunbauer, H.J.M. Rheological behaviour of hyperbranched polymer/montmorillonite clay nanocomposites. J. Rheol. 2004, 48, 1049–1065. [Google Scholar] [CrossRef]
  173. Lingaiah, S.; Shivakumar, K.N.; Sadler, R.; Sharpe, M. A method of visualization of dispersion of nanoplatelets in nanocomposites. Compos. Sci. Technol. 2005, 65, 2276–2280. [Google Scholar] [CrossRef]
  174. Azzouz, A.; Nistor, D.; Miron, D.; Ursu, A.-V.; Sajin, T.; Monette, F.; Niquette, P.; Hausler, R. Assessment of the acid–base strength distribution of ion-exchanged montmorillonites through NH3 and CO2-TPD measurements. Thermochim. Acta 2006, 449, 27–34. [Google Scholar] [CrossRef]
  175. Liu, P. Hyperbranched aliphatic polyester grafted attapulgite via a melt polycondensation process. Appl. Clay Sci. 2007, 35, 11–16. [Google Scholar] [CrossRef]
  176. Azzouz, A.; Assaad, E.; Ursu, A.-V.; Sajin, T.; Nistor, D.; Roy, R. Carbon dioxide retention over montmorillonite-dendrimer materials. Appl. Clay Sci. 2010, 48, 133–137. [Google Scholar] [CrossRef]
  177. Irani, M.; Fan, M.; Ismail, H.; Tuwati, A.; Dutcher, B.; Russell, A.G. Modified nanosepiolite as an inexpensive support of tetraethylenepentamine for CO2 sorption. Nano Energy 2015, 15, 235–246. [Google Scholar] [CrossRef]
  178. Yuan, M.; Gao, G.; Hu, X.; Luo, X.; Huang, Y.; Jin, B.; Liang, Z. Pre-modified sepiolite functionalized with triethylenetetramine as an effective and inexpensive adsorbent for CO2 capture. Ind. Eng. Chem. Res. 2018, 57, 6189–6200. [Google Scholar] [CrossRef]
  179. Thomas, J.; Bohor, B.F. Surface area of montmorillonite from the dynamic sorption of nitrogen and carbon dioxide. Clays Clay Miner. 1968, 16, 83–91. [Google Scholar] [CrossRef]
  180. Fripiat, J.J.; Cruz, M.I.; Bohor, B.F.; Thomas, J. Interlamellar adsorption of carbon dioxide by smectites. Clays Clay Miner. 1974, 22, 23–30. [Google Scholar] [CrossRef]
  181. Chan, W.H.; Mazlee, M.N.; Ahmad, Z.A.; Ishak, M.A.M.; Shamsul, J.B. The development of low-cost adsorbents from clay and waste materials: A review. J. Mater. Cycles Waste Manag. 2017, 19, 293–301. [Google Scholar] [CrossRef]
  182. Jeon, P.R.; Choi, J.; Yun, T.S.; Lee, C.H. Sorption equilibrium and kinetics of CO2 on clay minerals from subcritical to supercritical conditions: CO2 sequestration at nanoscale interfaces. Chem. Eng. J. 2014, 255, 705–715. [Google Scholar] [CrossRef]
  183. Giesting, P.; Guggenheim, S.; Koster van Groos, A.F.; Busch, A. Interaction of carbon dioxide with Na-exchanged montmorillonite at pressures to 640 bars: Implications for CO2 sequestration. Int. J. Greenh. Gas Control 2012, 8, 73–81. [Google Scholar] [CrossRef]
  184. Schaef, H.T.; Ilton, E.S.; Qafoku, O.; Martin, P.F.; Felmy, A.R.; Rosso, K.M. In situ XRD study of Ca2+ saturated montmorillonite (STX-1) exposed to anhydrous and wet supercritical carbon dioxide. Int. J. Greenh. Gas Control 2012, 6, 220–229. [Google Scholar] [CrossRef]
  185. Rother, G.; Ilton, E.S.; Wallacher, D.; HauB, T.; Schaef, H.T.; Qafoku, O.; Rosso, K.M.; Felmy, A.R.; Krukowski, E.G.; Stack, A.G.; et al. CO2 sorption to sub-single hydration layer montmorillonite clay studied by excess sorption and neutron diffraction measurements. Env. Sci. Technol. 2013, 47, 205–211. [Google Scholar] [CrossRef]
  186. Busch, A.; Alles, S.; Gensterblum, Y.; Prinz, D.; Dewhurst, D.N.; Raven, M.D.; Stanjek, H.; Krooss, B.M. Carbon dioxide storage potential of shales. Int. J. Greenh. Gas Control 2008, 2, 297–308. [Google Scholar] [CrossRef]
  187. Maruthi Sena, M.; Krishnan, M. Role of cations in adsorption of supercritical carbon dioxide at smectite mineral-water interfaces: Molecular dynamics and adaptive biasing fore simulation studies. J. Phys. Chem. C. 2018, 123, 1170–1184. [Google Scholar] [CrossRef]
  188. Narayanan Nair, A.K.; Cui, R.; Sun, S. Overview of the adsorption and transport properties of water, ions, carbon dioxide, and methane in swelling clays. ACS Earth Space Chem. 2021, 5, 2599–2611. [Google Scholar] [CrossRef]
  189. Mendel, N.; Sîretanu, D.; Sîretanu, I.; Brilman, D.W.; Mugele, F. Interlayer Cation-Controlled Adsorption of Carbon Dioxide in Anhydrous Montmorillonite Clay. J. Phys. Chem. C 2021, 125, 27159–27169. [Google Scholar] [CrossRef]
  190. Sato, K.; Hunger, M. Carbon dioxide adsorption in open nanospaces formed by overlap of saponite clay nanosheets. Comm. Chem. 2020, 3, 91. [Google Scholar] [CrossRef]
  191. Delgado, J.A.; Uguina, M.A.; Sotelo, J.L.; Ruı’z, B.; Rosa’rio, M. Carbon dioxide/methane separation by adsorption on sepiolite. J. Nat. Gas. Chem. 2007, 16, 235–243. [Google Scholar] [CrossRef]
  192. Pires, J.; Bestilleiro, M.; Pinto, M.; Gil, A. Selective adsorption of carbon dioxide, methane, and ethane by porous clays heterostructures. Sep. Purif. Technol. 2008, 61, 161–167. [Google Scholar] [CrossRef]
  193. Franco, F.; Cecilia, J.A.; Pozo, M.; Pardo, L.; Bellido, E.; Garcia-Sancho, C. Microwave assisted acid treatment of kerolitic clays from the Neogene Madrid Basin (Spain) and its use in CO2 capture processes. Microporous Mesoporous Mater. 2020, 292, 109749. [Google Scholar] [CrossRef]
  194. da Silva, G.J.; Fossum, J.O.; DiMasi, E.; Maloy, K.J. Hydration transitions in a nanolayered synthetic silicate: A synchrotron x-ray scattering study. Phys. Rev. B 2003, 67, 094114. [Google Scholar] [CrossRef]
  195. Jansson, M.; Eriksen, T.E. In situ anion diffusion experiments using radiotracers. J. Contam. Hydrol. 2004, 68, 183–192. [Google Scholar] [CrossRef]
  196. Tambach, T.J.; Hensen, E.J.M.; Smit, B. Molecular simulations of swelling clay minerals. J. Phys. Chem. B 2004, 108, 7586–7596. [Google Scholar] [CrossRef] [Green Version]
  197. Malikova, N.; Cadene, A.; Dubois, E.; Marry, C.; Durand-Vidal, S.; Turq, P.; Breu, J.; Longeville, S.; Zanotti, J.-M. Water diffusion in a synthetic hectorite clay studied by quasi-elastic neutron scattering. J. Phys. Chem. C 2007, 111, 17603–17611. [Google Scholar] [CrossRef]
  198. Porion, P.; Michot, L.J.; Faugere, A.M.; Delville, A. Structural and dynamical properties of the water molecules confined in dense clay sediments: A study combining 2H NMR spectroscopy and multiscale numerical modelling. J. Phys. Chem. C 2007, 111, 5441–5453. [Google Scholar] [CrossRef]
  199. Tenorio, R.P.; Alme, L.R.; Engelsberg, M.; Fossum, J.O.; Hallwass, F. Geometry and dynamics of intercalated water in Na-fluorhectorite clay hydrates. J. Phys. Chem. C 2008, 112, 575–580. [Google Scholar] [CrossRef]
  200. Tenorio, R.P.; Engelsberg, M.; Fossum, J.O.; da Silva, G.J. Intercalated water in synthetic fluorhectorite clay. Langmuir 2010, 26, 9703–9709. [Google Scholar] [CrossRef]
  201. Jimenez-Ruiz, M.; Ferrage, E.; Delville, A.; Michot, L.J. Anisotropy on the collective dynamics of water confined in swelling clay minerals. J. Phys. Chem. A 2012, 116, 2379–2387. [Google Scholar] [CrossRef]
  202. Franco, F.; Pozo, M.; Cecilia, J.A.; Benitez-Guerrero, M.; Pozo, E.; Rubi, J.A.M. Microwave assisted acid treatment of sepiolite: The role of composition and “crystallinity”. Appl. Clay Sci. 2014, 102, 15–27. [Google Scholar] [CrossRef]
  203. Silva, V.C.; Araujo, M.E.B.; Rodrigues, A.M.; Cartazo, J.M.; Menezes, R.R.; Neves, G.A. adsorption behaviour of acid-treated Brazilian palygorskite for cationic and anionic dyes removal from water. Sustainability 2021, 13, 3954. [Google Scholar] [CrossRef]
  204. Wal, K.; Rutkowski, P.; Stawinski, W. Application of clay minerals and their derivatives in adsorption from gaseous phase. Appl. Clay Sci. 2021, 215, 106323. [Google Scholar] [CrossRef]
  205. Chen, Y.H.; Lu, D.L. Amine modification on kaolinites to enhance CO2 adsorption. J. Colloid Interface Sci. 2014, 436, 47–51. [Google Scholar] [CrossRef]
  206. Horri, N.; Sanz-Pérez, E.S.; Arencibia, A.; Sanz, R.; Frini-Srasra, N. Amine grafting of acid-activated bentonite for carbon dioxide capture. Appl. Clay Sci. 2019, 180, 105195. [Google Scholar] [CrossRef]
  207. Ramadass, K.; Sathish, C.I.; Singh, G.; Ruban, S.J.; Ruban, A.M.; Bahadur, R.; Vinu, A. Morphologically tuneable nano architectonics of mixed kaolin-halloysite derived nitrogen-doped activated nanoporous carbons for supercapacitor and CO2 capture applications. Carbon 2022, 192, 133–144. [Google Scholar] [CrossRef]
  208. Bo Hunvik, K.W.; Loch, P.; Wallacher, D.; Kirch, A.; Cavalcanti, L.P.; Riess, M.; Fossum, J.O. CO2 adsorption enhanced by tuning the layer charge in a clay mineral. Langmuir 2021, 37, 14491–14499. [Google Scholar] [CrossRef]
  209. Niu, M.; Yang, H.; Zhang, X.; Wang, Y.; Tang, A. Amine-impregnated mesoporous silica nanotube as an emerging nanocomposite for CO2 capture. ACS Appl. Mater. Interface 2016, 8, 17312–17320. [Google Scholar] [CrossRef]
  210. Roth, E.A.; Agarwal, S.; Gupta, R.K. Nanoclay-based solid sorbents for CO2. Energy Fuels 2013, 27, 4129–4136. [Google Scholar] [CrossRef]
  211. Pinto, M.L.; Mafra, L.; Guil, J.M.; Pires, J.; Rocha, J. Adsorption and activation of CO2 by amine-modified nanoporous materials studied by solid-state NMR and 13CO2 adsorption. Chem. Mater. 2011, 23, 1387–1395. [Google Scholar] [CrossRef]
  212. Pinto, M.L.; Pires, J. Porous and hybrid clay-based materials for separation of hydrocarbons. Microporous Mesoporous Mater. 2012, 151, 403–410. [Google Scholar] [CrossRef]
  213. Garea, S.A.; Mihai, A.J.; Ghebaur, A.; Nistor, C.; Sarbu, A. Porous clay heterostructures: A new inorganic host for 5-fluorouracil encapsulation. Int. J. Pharm. 2015, 492, 299–309. [Google Scholar] [CrossRef] [PubMed]
  214. Chmielarz, L.; Piwowarska, Z.; Kustrowski, P.; Wegrzyn, A.; Gil, B.; Kowalczyk, A.; Dudek, B.; Dziembaj, R.; Michalik, M. Comparison study of titania pillared interlayered clay and porous clay heterostructures modified with copper and iron as catalysts of the DeNOx process. Appl. Clay Sci. 2011, 53, 164–173. [Google Scholar] [CrossRef]
  215. Cecilia, J.A.; Arango-Diaz, A.; Jimenez-Jimenez, J.; Storaro, L.; Moretti, E.; Rodriguez-Castellon, E. CuO-CeO2 supported on montmorillonite-derived porous clay heterostructures (PCH) for preferential CO oxidation in H2-rich stream. Catal. Today 2015, 253, 126–136. [Google Scholar] [CrossRef]
  216. Soriano, M.D.; Cecilia, J.A.; Natoli, A.; Jimenez-Jimenez, J.; Nieto, J.M.L.; Rodriguez-Castellon, E. Vanadium oxide supported on porous clay heterostructure for the partial oxidation of hydrogen sulphide to sulphur. Catal. Today 2015, 254, 36–42. [Google Scholar] [CrossRef]
  217. Saboya, R.M.A.; Cecilia, J.A.; Garcia-Sancho, C.; Luna, F.M.T.; Rodriguez-Castellon, E.; Cavalcante, C.L. WO3-based catalysts supported on porous clay heterostructures (PCH) with Si–Zr pillars for synthetic esters production. Appl. Clay Sci. 2016, 124–125, 69–78. [Google Scholar] [CrossRef]
  218. Azzouz, A.; Platon, N.; Nousir, S.; Ghomari, K.; Nistor, D.; Shiao, T.C.; Roy, R. OH-enriched organo montmorillonites for potential applications in carbon dioxide separation and concentration. Sep. Purif. Technol. 2013, 108, 181–188. [Google Scholar] [CrossRef]
  219. Cavenati, S.; Grande, C.A.; Rodrigues, A.E. Separation of CH4/CO2/N2 mixtures by layered pressure swing adsorption for upgrade of natural gas. Chem. Eng. Sci. 2006, 61, 3893–3906. [Google Scholar] [CrossRef]
  220. Grande, C.A.; Rodrigues, A.E. Electric swing adsorption for CO2 removal from flue gas. Int. J. Greenh. Gas Control 2008, 2, 194–202. [Google Scholar] [CrossRef]
  221. Zhang, J.; Webley, P.A.; Xiao, P. Effect of process parameters on power requirements of vacuum swing adsorption technology for CO2 capture from flue gas. Energy Convers. Manag. 2008, 49, 346–356. [Google Scholar] [CrossRef]
  222. Gupta, S.K.; Lesslie, R.D.; King, A.D. Solubility of alcohols in compressed gases. A comparison of vapor-phase interactions of alcohols and homomorphic compounds with various gases I. Ethanol in compressed helium, hydrogen, argon, methane, ethylene, ethane, carbon dioxide, and nitrous oxide. J. Phys. Chem. 1973, 77, 2011–2015. [Google Scholar] [CrossRef]
  223. Massoudi, R.; King, A.D. Solubility of alcohols in compressed gases. A comparison of vapor-phase interactions of alcohols and homomorphic compounds with various gases II. 1-Butanol, diethyl ether, and n-pentane in compressed nitrogen, argon, methane, ethane, and carbon dioxide at 25.deg. J. Phys. Chem. 1973, 77, 2016–2018. [Google Scholar] [CrossRef]
  224. Saharay, M.; Balasubramanian, S. Electron donor–acceptor interactions in ethanol—CO2 mixtures: An ab initio molecular dynamics study of supercritical carbon dioxide. J. Phys. Chem. B 2006, 110, 3782–3790. [Google Scholar] [CrossRef] [PubMed]
  225. Azzouz, A.; Ursu, A.-V.; Nistor, D.; Sajin, T.; Assaad, E.; Roy, R. TPD study of the reversible retention of carbon dioxide over montmorillonite intercalated with polyol dendrimers. Thermochim. Acta 2009, 496, 45–49. [Google Scholar] [CrossRef]
  226. Calderon, M.; Quadir, M.A.; Sharma, S.K.; Haag, R. Dendritic polyglycerols for biomedical applications. Adv. Mater. 2010, 22, 190–218. [Google Scholar] [CrossRef]
  227. Gassensmith, J.J.; Furukawa, H.; Smaldone, R.A.; Forgan, R.S.; Botros, Y.Y.; Yaghi, O.M.; Stoddart, J.F. Strong and reversible binding of carbon dioxide in a green metal–organic framework. J. Am. Chem. Soc. 2011, 133, 15312–15315. [Google Scholar] [CrossRef]
  228. Garcia-Gallastegui, A.; Iruretagoyena, D.; Gouvea, V.; Mokhtar, M.; Asiri, A.M.; Basahel, S.N.; Al-Thabaiti, S.A.; Alyoubi, A.O.; Chadwick, D.; Sha, M.S.P. Graphene oxide as support for layered double hydroxides: Enhancing the CO2 adsorption capacity. Chem. Mater. 2012, 24, 4531–4539. [Google Scholar] [CrossRef]
  229. Galarneau, A.; Barodawalla, A.; Pinnavaia, T.J. Porous clay heterostructures formed by gallery-templated synthesis. Nature 1995, 374, 529–531. [Google Scholar] [CrossRef]
  230. Cecilia, J.A.; Garcia-Sancho, C.; Franco, F. Montmorillonite based porous clay heterostructures: Influence of Zr in the structure and acidic properties. Microporous Mesoporous Mater. 2013, 176, 95–102. [Google Scholar] [CrossRef]
  231. Humphrey, J.; Boyd, D. Clay: Types, Properties and Uses; Nova Science Publishers, Inc.: New York, NY, USA, 2011. [Google Scholar]
  232. Nawani, P.; Gelfer, M.Y.; Hsiao, B.S.; Frenkel, A.; Gilman, J.W.; Khalid, S. Surface Modification of nanoclays by catalytically active transition metal Ions. Langmuir 2007, 23, 9808–9815. [Google Scholar] [CrossRef]
  233. Zope, I.S.; Dasari, A.; Camino, G. Elucidating the catalytic effect of metal ions in montmorillonite on thermal degradation of organic modifier. Mater. Chem. Phys. 2015, 157, 69–79. [Google Scholar] [CrossRef]
  234. Reddy, C.R.; Bhat, Y.S.; Nagendrappa, G.; Jai Prakash, B.S. Brønsted and Lewis acidity of modified montmorillonite clay catalysts determined by FT-IR spectroscopy. Catal. Today 2009, 141, 157–160. [Google Scholar] [CrossRef] [Green Version]
  235. Tang, T.; Chen, X.; Meng, X.; Chen, H.; Ding, Y. Synthesis of multiwalled carbon nanotubes by catalytic combustion of polypropylene. Angew. Chem. Int. Ed. 2005, 44, 1517–1520. [Google Scholar] [CrossRef] [PubMed]
  236. Tang, T.; Chen, X.; Chen, H.; Meng, X.; Jiang, Z.; Bi, W. Catalyzing carbonization of polypropylene itself by supported nickel catalyst during combustion of polypropylene/clay nanocomposite for improving fire retardancy. Chem. Mater. 2005, 17, 2799–2802. [Google Scholar] [CrossRef]
  237. Hu, Y.; Wang, X.; Li, J. Regulating effect of exfoliated clay on intumescent char structure and flame retardancy of polypropylene composites. Ind. Eng. Chem. Res. 2016, 55, 5892–5901. [Google Scholar] [CrossRef]
  238. Vincente, M.A.; Gil, A.; Bergaya, F. Pillared clays and clay minerals. Dev. Clay. Sci. 2013, 5, 523–557. [Google Scholar] [CrossRef]
  239. Gil, A.; Vicente, M.A. Progress and perspectives on pillared clays applied in energetic and environmental remediation processes. Curr. Opin. Green Sustain. Chem. 2020, 21, 56–63. [Google Scholar] [CrossRef]
  240. Sayari, A.; Belmabkhout, Y.; Serna-Guerrero, R. Flue gas treatment via CO2 adsorption. Chem. Eng. J. 2011, 171, 760–774. [Google Scholar] [CrossRef]
  241. Sanz, R.; Calleja, G.; Arencibia, A.; Sanz-Perez, E.S. CO2 adsorption on branched polyethyleneimine-impregnated mesoporous silica SBA-15. Appl. Surf. Sci. 2010, 256, 5323–5328. [Google Scholar] [CrossRef]
  242. Fernandes, A.C.; Pinto, M.L.; Antunes, F.; Pires, J. L-histidine-based organoclays for the storage and release of therapeutic nitric oxide. J. Mater. Chem. B. 2015, 3, 3556–3563. [Google Scholar] [CrossRef]
  243. Elkhalifah, A.E.I.; Bustam, M.A.; Azmi, M.S.; Murugesan, T. Carbon dioxide retention on bentonite clay adsorbents modified by mono-, di- and triethanolamine compounds. Adv. Mater. Res. 2014, 917, 115–122. [Google Scholar] [CrossRef]
  244. Das, S.; Maity, A.; Pradhan, M.; Jana, S. Assessing atmospheric CO2 entrapped in clay nanotubes using residual gas analyser. Anal. Chem. 2016, 88, 2205–2211. [Google Scholar] [CrossRef] [PubMed]
  245. Jana, S.; Das, S.; Ghosh, C.; Maity, A.; Pradhan, M. Halloysite nanotubes capturing isotope selective atmospheric CO2. Sci. Rep. 2015, 5, 8711. [Google Scholar] [CrossRef] [PubMed]
  246. Zhang, Y.; Chen, M.; Li, G.; Shi, C.; Wang, B.; Ling, Z. Exfoliated vermiculite nanosheets supporting tetraethylenepentamine for CO2 capture. Results Mater. 2020, 7, 100102. [Google Scholar] [CrossRef]
  247. Atilhan, M.; Atilhan, S.; Ullah, R.; Anaya, B.; Cagin, T.; Yavuz, C.T.; Aparicio, S. High-pressure methane, carbon dioxide and nitrogen adsorption on amine-impregnated porous montmorillonite nano clays. J. Chem. Eng. Data. 2016, 61, 2749–2760. [Google Scholar] [CrossRef]
  248. Ouyang, J.; Zheng, C.; Gu, W.; Zhang, Y.; Yang, H.; Suib, S.L. Textural properties determined CO2 capture of Tetraethylenepentamine loaded SiO2 nanowires from α-sepiolite. Chem. Eng. J. 2018, 337, 342–350. [Google Scholar] [CrossRef]
  249. Ouyang, J.; Gu, W.; Zheng, C.; Yang, H.; Zhang, X.; Jin, Y.; Chen, J.; Jiang, J. Polyethyleneimine (PEI) loaded MgO-SiO2 nanofibers from sepiolite minerals for reusable CO2 capture/release applications. Appl. Clay Sci. 2018, 152, 267–275. [Google Scholar] [CrossRef]
  250. Liu, L.; Chen, H.; Shiko, E.; Fan, X.; Zhou, Y.; Zhang, G.; Luo, X.; Hu, X.E. Low-cost DETA impregnation of acid-activated sepiolite for CO2 capture. Chem. Eng. J. 2018, 353, 940–948. [Google Scholar] [CrossRef]
  251. Stevens, L.; Williams, K.; Han, W.Y.; Drage, T.; Snape, C.; Wood, J.; Wang, J. Preparation and CO2 adsorption of diamine modified montmorillonite via exfoliation grafting route. Chem. Eng. J. 2013, 215, 699–708. [Google Scholar] [CrossRef]
  252. Elkhalifah, A.E.I.; Bustam, M.A.; Shariff, A.M.; Murugesan, T. Selective adsorption of CO2 on a regenerable amine-bentonite hybrid adsorbent. Appl. Clay Sci. 2015, 107, 213–219. [Google Scholar] [CrossRef]
  253. Pires, J.; Juźków, J.; Pinto, M.L. Amino acid modified montmorillonite clays as sustainable materials for carbon dioxide adsorption and separation. Colloids Surf. A Physicochem. Eng. 2018, 544, 105–110. [Google Scholar] [CrossRef]
  254. Chen, C.; Park, D.W.; Ahn, W.S. Surface modification of a low-cost bentonite for post-combustion CO2 capture. Appl. Surf. Sci. 2013, 283, 699–704. [Google Scholar] [CrossRef]
  255. Horsch, S.; Serhatkulu, G.; Gulari, E.; Kannan, R.M. Supercritical CO2 dispersion of nano-clays and clay/polymer nanocomposites. Polymer 2006, 47, 7485–7496. [Google Scholar] [CrossRef]
  256. Gao, F. Clay/polymer composites: The story. Mater. Today 2004, 7, 50–55. [Google Scholar] [CrossRef]
  257. Ismail, N.M.; Ismail, A.F.; Mustafa, A.; Zulhairun, A.K.; Aziz, F.; Bolong, N.; Razali, A.R. Polymer clay nanocomposites for gas separation: A review. ECR 2019, 2, 1–5. [Google Scholar] [CrossRef]
  258. Petersell, V.; Juriado, K.; Raukas, A.; Shtokalenko, M.; That-Kok, K. Quaternary deposits and weathered bedrock material as a source of dangerous radon emissions in Estonia. Geologos 2015, 21, 139–147. [Google Scholar] [CrossRef] [Green Version]
  259. Tuuling, I. The Leba Ridge-Riga-Pskov fault zone—A major East European Craton interior dislocation zone and its role in the early Palaeozoic development of the platform cover. Est. J. Earth Sci. 2019, 68, 161–189. [Google Scholar] [CrossRef]
  260. Raukas, A.; Teedumae, A. Geology of Mineral Resources of Estonia; Estonian Academy Publishers: Tallin, Estonia, 1997. [Google Scholar]
  261. Geology and Mineral Resources of Estonia. Available online: https://geoloogia.info/geology/text.html (accessed on 7 November 2021).
  262. Gorokhov, I.M.; Melnikov, N.N.; Kuznetsov, A.B.; Konstantinova, G.V.; Turchenko, T.L. Sm-Md systematics of fine-grained fractions of the Lower Cambrian blue clay in northern Estonia. Lithol. Miner. Resourc. 2007, 42, 482–495. [Google Scholar] [CrossRef]
  263. Lescinskis, O.; Svinka, R.; Svinka, V. Characterization of three different clay minerals from different geological periods. In Proceedings of the 2nd International Symposium “Clays and Ceramics”, Riga, Latvia, 29–31 January 2018; Dzene, L., Vircava, I., Eds.; University of Latvia: Riga, Latvia, 2018; p. 26. [Google Scholar]
  264. Stinkulis, G. Middle Devonian clays of the Liepa clay deposit, Latvia: Composition, sedimentary environment and deformation structures. In Proceedings of the 2nd International Symposium “Clays and Ceramics”, Riga, Latvia, 29–31 January 2018; Dzene, L., Vircava, I., Eds.; University of Latvia: Riga, Latvia, 2018; pp. 16–17. [Google Scholar]
  265. Burlakovs, J.; Ozola, R.; Kostjukovs, J.; Klavins, I.; Purmalis, O.; Klavins, M. Properties of the Jurassic clayey deposits of southwestern Latvia and northern Lithuania. Mater. Sci. Appl. Chem. 2015, 32, 5–12. [Google Scholar] [CrossRef] [Green Version]
  266. Hamza, A.; Hussein, I.A.; Al-Marri, M.J.; Mahmoud, M.; Shawabkeh, R. Impact of clays on CO2 adsorption and enhanced gas recovery in sandstone reservoirs. Int. J. Greenh. Gas Control 2021, 106, 103286. [Google Scholar] [CrossRef]
  267. Hu, X.; Deng, H.; Lu, C.; Tian, Y.; Jin, Z. Characterization of CO2/CH4 competitive adsorption in various clay minerals in relation to shale gas recovery from molecular simulation. Energy Fuels 2019, 33, 8202–8214. [Google Scholar] [CrossRef]
  268. Zhou, C.H.; Shen, Z.-F.; Liu, L.-H.; Liu, S.-M. Preparation and functionality of clay-containing films. J. Mater. Chem. 2011, 21, 15132–15153. [Google Scholar] [CrossRef]
  269. Bergaya, F.; Lagaly, G. Handbook of Clay Science. Developments in Clay Science 5, 2nd ed.; Elsevier: Amsterdam, The Netherlands, 2013. [Google Scholar]
  270. Tahir, M.; Saidina, N. Photocatalytic CO2 reduction by CH4 over montmorillonite modified TiO2 nanocomposites in a continuous monolith photoreactor. Appl. Catal. B-Environ. 2013, 142–143, 512–523. [Google Scholar] [CrossRef]
  271. Christine, S.; Mana, A.; Hanafiah, M.M.; Khan, A.J. Environmental characteristics of clay and clay-based minerals. Geol. Ecol. Landsc. 2017, 1, 155–161. [Google Scholar] [CrossRef] [Green Version]
  272. Galan, E.; Aparicio, P.; Miras, A. Sepiolite and palygorskite as sealing materials for the geological storage of carbon dioxide. Dev. Clay. Sci. 2011, 3, 375–392. [Google Scholar] [CrossRef]
  273. Loganathan, N.; Yazaydin, A.O.; Bowers, G.M.; Ngouana-Wakou, B.F.; Kalinichev, A.G.; Kirkpatrick, R.J. Role of cations in the methane/carbon dioxide partitioning in nano- and mesopores of illite using constant reservoir composition molecular dynamics simulation. J. Phys. Chem. C 2020, 124, 2490–2500. [Google Scholar] [CrossRef]
  274. Zhang, J.; Clennell, M.B.; Liu, K.; Pervukhina, M.; Chen, G.; Dewhurst, D.N. Methane and carbon dioxide adsorption on illite. Energy Fuels 2016, 30, 10643–10652. [Google Scholar] [CrossRef]
  275. Jin, Z.; Firoozabadi, A. Methane and carbon dioxide adsorption in clay-like slit pores by Monte Carlo simulations. Fluid Phase Equilib. 2013, 360, 456–465. [Google Scholar] [CrossRef]
  276. 10 Carbon Capture Methods Compared: Costs, Scalability, Permanence, Cleanness. Available online: https://energypost.eu/10-carbon-capture-methods-compared-costs-scalability-permanence-cleanness/ (accessed on 16 November 2021).
  277. Rosenbloom, D.; Markard, J.; Geels, F.W.; Fuenfschilling, L. Opinion: Why carbon pricing is not sufficient to mitigate climate change—and how “sustainability transition policy” can help. Proc. Natl. Acad. Sci. USA 2020, 117, 8664–8668. [Google Scholar] [CrossRef] [Green Version]
  278. Stiglitz, J.E.; Stern, N.; Duan, M.; Edenhofer, O.; Giraud, G.; Heal, G.M.; Ia Rovere, E.L.; Morris, A.; Moyer, E.; Pangestu, M.; et al. Report of the High-Level Commission on Carbon Prices; International Bank for Reconstruction and Development and International Development Association/The World Bank: Washington, DC, USA, 2017. [Google Scholar]
  279. Nordhaus, W.D. The Climate Casino: Risk, Uncertainty and Economics for a Warming World; Yale University Press: London, UK, 2013. [Google Scholar]
  280. Markard, J.; Raven, R.; Truffer, B. Sustainability transitions: An emerging field of research and its prospects. Res. Policy 2012, 41, 955–967. [Google Scholar] [CrossRef]
Figure 1. A schematic representation of silica tetrahedron and magnesium or aluminium octahedron.
Figure 1. A schematic representation of silica tetrahedron and magnesium or aluminium octahedron.
Minerals 12 00349 g001
Figure 2. Structural schemes for some of the clay minerals: (a) kaolinite; (b) halloysite; (c) sepiolite; (d) vermiculite, and (e) smectite (adapted from Schulze 2005) [141].
Figure 2. Structural schemes for some of the clay minerals: (a) kaolinite; (b) halloysite; (c) sepiolite; (d) vermiculite, and (e) smectite (adapted from Schulze 2005) [141].
Minerals 12 00349 g002
Table 1. Mean annual CO2 growth rates (ppm y−1) for Mauna Loa for the last decade from 2011 through 2021 [10].
Table 1. Mean annual CO2 growth rates (ppm y−1) for Mauna Loa for the last decade from 2011 through 2021 [10].
Year20112012201320142015201620172018201920202021
Growth (ppm)1.922.651.992.222.903.031.922.862.482.312.46 1
1 Data for October 2021.
Table 2. Share of electricity production (%) in the Baltic States in 2020. Fossil fuel, low-carbon sources (solar, wind, hydropower, biomass, waste, geothermal, wave, tidal) [38].
Table 2. Share of electricity production (%) in the Baltic States in 2020. Fossil fuel, low-carbon sources (solar, wind, hydropower, biomass, waste, geothermal, wave, tidal) [38].
ShareEstoniaLatviaLithuania
Fossil fuel64.7938.8942.46
Low-carbon sources (renewables)35.2161.1157.54
Table 3. Energy consumption by source (TWh) in the Baltic States in 2019 [38].
Table 3. Energy consumption by source (TWh) in the Baltic States in 2019 [38].
ShareEstoniaLatviaLithuania
Oil182238
Coal39<12
Gas41322
Hydropower<15<1
Wind2<14
Solar<10<1
Other renewables321
Table 4. Comparison of methods for CO2 reduction. Costs and scalability [276].
Table 4. Comparison of methods for CO2 reduction. Costs and scalability [276].
ApproachCosts
(Euro/tCO2)
Scalability
(GtCO2/y)
Reducing CO2 to its constituent components−70 to 2600.3–0.6
Hydrocarbon fuel production5901.0–4.2
CO2 fixation with microalgae200 to 8000.2–0.9
Use in building materials−25 to 600.1–1.4
Enhanced oil recovery−35 to 500.1–1.8
Bioenergy with carbon capture50 to 1400.5–5.0
Forestry−35 to 101.5
Soil carbon sequestration−80 to 200.9–1.9
Biochar−600.2–1.0
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Krūmiņš, J.; Kļaviņš, M.; Ozola-Davidāne, R.; Ansone-Bērtiņa, L. The Prospects of Clay Minerals from the Baltic States for Industrial-Scale Carbon Capture: A Review. Minerals 2022, 12, 349. https://doi.org/10.3390/min12030349

AMA Style

Krūmiņš J, Kļaviņš M, Ozola-Davidāne R, Ansone-Bērtiņa L. The Prospects of Clay Minerals from the Baltic States for Industrial-Scale Carbon Capture: A Review. Minerals. 2022; 12(3):349. https://doi.org/10.3390/min12030349

Chicago/Turabian Style

Krūmiņš, Jānis, Māris Kļaviņš, Rūta Ozola-Davidāne, and Linda Ansone-Bērtiņa. 2022. "The Prospects of Clay Minerals from the Baltic States for Industrial-Scale Carbon Capture: A Review" Minerals 12, no. 3: 349. https://doi.org/10.3390/min12030349

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