Next Article in Journal
Evaluation of the Use of Different Dedicated Mechanical Subcooling (DMS) Strategies in a Water Source Transcritical CO2 Heat Pump for Space Heating Applications
Previous Article in Journal
Steam Explosion Pre-Treatment of Sawdust for Biofuel Pellets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Carbon Capture and Use (CCU) Technologies: A Comprehensive Review and CO2 Mitigation Potential Analysis

by
Christiano B. Peres
1,
Pedro M. R. Resende
2,3,4,
Leonel J. R. Nunes
2,5,* and
Leandro C. de Morais
1
1
Institute of Science and Technology, São Paulo State University (UNESP) “Júlio de Mesquita Filho”, Sorocaba Campus, Av. Três de Março, 511, Alto da Boa Vista, Sorocaba 18087-180, São Paulo, Brazil
2
Prometheus, Instituto Politécnico de Viana do Castelo, Rua da Escola Industrial e Comercial de Nun’Alvares, 4900-347 Viana do Castelo, Portugal
3
Escola Superior de Tecnologia e Gestão, Instituto Politécnico de Viana do Castelo, Avenida do Atlântico, n.º 644, 4900-348 Viana do Castelo, Portugal
4
CEFT, Faculdade de Engenharia, Universidade do Porto, 4200-465 Porto, Portugal
5
Escola Superior Agrária, Instituto Politécnico de Viana do Castelo, Rua D. Mendo Afonso, 147, Refóios do Lima, 4990-706 Ponte de Lima, Portugal
*
Author to whom correspondence should be addressed.
Clean Technol. 2022, 4(4), 1193-1207; https://doi.org/10.3390/cleantechnol4040073
Submission received: 1 July 2022 / Revised: 26 September 2022 / Accepted: 8 November 2022 / Published: 17 November 2022

Abstract

:
One of society’s major current challenges is carbon dioxide emissions and their consequences. In this context, new technologies for carbon dioxide (CO2) capture have attracted much attention. One of these is carbon capture and utilization (CCU). This work focuses on the latest trends in a holistic approach to carbon dioxide capture and utilization. Absorption, adsorption, membranes, and chemical looping are considered for CO2 capture. Each CO2 capture technology is described, and its benefits and drawbacks are discussed. For the use of carbon dioxide, various possible applications of CCU are described, starting with the utilization of carbon dioxide in agriculture and proceeding to the conversion of CO2 into fuels (catalytic processes), chemicals (photocatalytic processes), polymers, and building supplies. For decades, carbon dioxide has been used in industrial processes, such as CO2-enhanced oil recovery, the food industry, organic compound production (such as urea), water treatment, and, therefore, the production of flame retardants and coolants. There also are several new CO2-utilization technologies at various stages of development and exploitation, such as electrochemical conversion to fuels, CO2-enhanced oil recovery, and supercritical CO2. At the end of this review, future opportunities are discussed regarding machine learning (ML) and life cycle assessment (LCA).

1. Introduction

Overreliance on fossil fuels has contributed to a critical energy crisis and to the increase in carbon dioxide in the atmosphere. CO2 is by far the most dangerous greenhouse gas per unit of concentration emitted by humans [1]. According to the International Energy Agency (IEA) [2], about 36 billion tons of carbon dioxide are emitted each year through transportation, power generation, cement manufacturing, deforestation, agriculture, and many other practices [2]. Most of the carbon dioxide emitted today will remain in the atmosphere for over 1000 years [3,4,5]. According to the National Ocean and Atmospheric Administration (NOAA) data, the average concentration of carbon dioxide in the atmosphere has increased dramatically, from 172–300 ppm before the last industrial era to 419 ppm in May 2021. In recent decades, extreme weather and signs of climate disaster, such as ocean acidification, air pollution, tsunamis, droughts, and species extinctions, have been seen as a result of climate warming and the effect of accumulation carbon dioxide emissions [6,7].
The Intergovernmental Panel on Climate Change (IPCC) has determined that the main causes of the increase in the carbon dioxide concentration are fossil fuel burning (mainly from industrial sources such as cement and power plants, iron manufacturing, and the chemical industry) and deforestation (agriculture). Human activity is responsible for most of the greenhouse gas (GHG) emissions. Carbon dioxide accounts for nearly 77% (or 35 billion tons) of greenhouse gas emissions, seriously affecting the environment and people’s way of life; this quantity of GHG can be associated with detrimental effects on human health, such as headaches and respiratory diseases [7]. Therefore, there is a crucial need to curb carbon dioxide emissions for the benefit of future generations. In this context, carbon dioxide capture and utilization (CCU) technology has gained considerable attention due to its important role in reducing greenhouse gas emissions [6,7,8].
Carbon capture and utilization (CCU) and carbon capture and storage (CCS) technologies play an increasingly important role since the purpose of their use is to promote the recycling of carbon dioxide (CO2) in the form of raw material—for example, in the replacement of fossil fuels, for the chemical industry, or even to produce new carbon-based materials. This use of CCU or CCS technologies can effectively contribute to creating net negative greenhouse gas emissions. Given growing interest in the subject, several works have been carried out in recent years. These include works explicitly addressing the issue of processes and technologies—that is, more pioneering works on the means of capturing and storing CO2. In addition, several works have focused on the technical and economic feasibility of these technologies, since, in the first analysis, it has been verified that all of them, without considering their ability to capture and store CO2, are intensive energy consumers and require significant initial investment for their implementation. In other words, these projects tend to require significant investment in the CAPEX and OPEX phases, which require complex feasibility analysis processes for correct decision making, as described in the recent work by Liu et al. [9]. These authors suggest that their decision-making model can be used regarding investment to be made in the technology learning effect process, which uses the component-based two-factor technology learning curve approach to predict the component costs. This process can also be implemented for enhanced oil recovery projects with CO2 (CO2-EOR), for which the price of a barrel of oil must reach USD 80. However, the same authors also concluded that, for methanol synthesis projects, the price per ton of methanol must reach USD 580 to justify the use of this technology. In other words, according to the authors, investment in CO2-MET projects is more economical than that in CO2-EOR projects. As seen in work by Liu et al. [9], in addition to Wich et al., Pires da Mota Costa et al., and Dimitriou et al. [10,11,12], after the technical validation of the methods to be used, the subsequent economic and financial validation assumes a decisive role in the feasibility of the projects.
Another relevant aspect is the analysis of the potential environmental impacts of these technologies, namely through a life-cycle assessment (LCA), as a way of guaranteeing that the CCU and CCS processes have environmental advantages over conventional production processes. Based on this perspective, authors such as Garcia-Garcia et al. [13] performed thorough and systematic environmental impact analyses. These authors conclude that, although the creation of net negative greenhouse gas emissions may initially seem to be a highly favorable aspect, the fact that these processes are energy-intensive means that, to achieve environmental improvements, renewable energy must be used, mainly producing hydrogen from water electrolysis. In other words, the use of these CCU and CCS technologies depends not only on the technological process of capturing and storing CO2, but also on the development of forms of renewable energy that can contribute to the efficiency of the process.
In recent years, significant progress has been made in increasing the availability of CCU materials (fuels, chemicals, and building materials that convert CO2), and their performance, to effectively reduce CO2 emissions. Given the lack of reviews of the progress of promising state-of-the-art CCU technologies, there is an urgent need for a full and timely summary of these advanced technologies and a comprehensive understanding of them. The purpose of this work is to outline and comprehensively describe recent progress in CCU. Research and/or review articles are surveyed to discover innovative aspects and technologies related to pre- and post-combustion CO2 capture, liquid and solid sorbents, carbon dioxide conversion via homogeneous and heterogeneous catalysis, and the direct use of carbon dioxide. First, we summarize the CCU scheme and discuss details of the processes involved. Next, we present a framework for the performance of liquid and solid CO2 capture materials under different operating conditions and propose possible scenarios to improve their carbon capture capacity and progress in CO2 utilization in different fields, systematically comparing and discussing their industrial applications. Finally, we present future perspectives and technological advances in the field.

2. CCU Technologies

2.1. Framework

CO2 capture options involve carbon capture and storage (CCS), which continually stores carbon dioxide and prevents it from entering the air, and carbon capture and utilization (CCU), which absorbs carbon dioxide in various industrial supply chains. CCS and CCU both aid a circular carbon economy, where carbon emissions are reduced, reused, recovered, and removed. Overall, CCU processes are critical to addressing global emissions and carbon-intensive manufacturing. The CCU scenario is considered an outstanding approach to combating anthropogenic carbon emissions [14,15,16]. Carbon capture is the most economical way to control industrial point-source carbon emissions [6,7,14]. Subsequent use refers to the use of captured carbon dioxide to synthesize value-added compounds or enhance oil recovery [6,7].
CCU concentrates on resource efficiency and aligns with the concept of a circular economy [17]: the capture and use of carbon dioxide provides the benefits of mitigating climate change and producing economically viable products [18]. The promising energy-saving techniques and their utilization in value-added merchandise are essential to the industry’s uptake of CCU. Despite research, technical and economic barriers limit most laboratory-scale technologies [16]. CCU comprises four steps: CO2 capture, CO2 utilization, CO2 storage, and CO2 transportation [6,19,20]. In this work, capture and utilization are described.

2.2. CO2 Capture

Carbon dioxide is emitted via burning fossil fuels by industrial sources, such as hydrogen power plants, cement manufacturers, coal gasification, power conversion, iron and steel plants, chemical production, and the petroleum industry [6,20]. Carbon capture methods are divided into two types: CO2 capture techniques and CO2 separation techniques [16].
CO2 capture technologies consist of the dissociation and storage of CO2 generated in industrial processes, or the removal of oxygen from the air, thereby decreasing the difficulty of capture and reducing the energetic demand [21]. Carbon capture technology is classified into three approaches: post-combustion capture, oxy-fuel combustion, and pre-combustion capture, of which post-combustion is a globally used approach. Pre-combustion separation is usually adopted to produce syngas (H2 and CO) through chemical reactions: CO and H2O are converted into CO2 and H2 through further conversion reactions, and then CO2 is separated out. This technique is largely utilized in natural gas boilers and coal-fired power plants [21,22]. By contrast, post-combustion techniques, such as absorption, adsorption, chemical looping, or membranes, isolate CO2 in the flue gas [16].
The oxygen-enriched combustion technique usually adopts pure oxygen or a mixture of air/carbon dioxide rather than oxygen for combustion. After denitrification and combustion, an elevated-purity CO2 can result for direct storage, which also has significant benefits compared with general combustion air. Oxygen-enriched fuel combustion produces CO2 and H2O, and about 75% less flue gas compared to oxy-fuel combustion [16,23]. Several physical and chemical methods are used in capture technology, containing solvent-based absorption, solid sorbents, membranes for CO2 capture, and chemical looping, as shown in Figure 1 [16,21,22,24].
Financially, carbon capture projects currently cost around 60–110 USD per ton globally and are expected to drop to around 30–50 USD per ton by 2030. This will encourage the rollout of the technology on a commercial scale [22]. As shown in Figure 2, the main CO2 separation techniques are absorption, adsorption, membranes, and chemical looping. The next section provides more information about these methods [24,25].

2.2.1. Absorption-Based CO2 Capture

In absorption methods for CO2 capture, a liquid absorbent (solvent) physically or chemically isolates CO2 from flue gas by forming a solution or compound (Figure 3). This technique has been used extensively both post-combustion (using chemical solvents) and pre-combustion (using a physical solvent) [5,16,26]. Chemisorption of amine-based solvents such as ammonia, monoethanolamine (MEA), diethanolamine (DEA), and N-methyldiethanolamine (MDEA), in addition to basic solvents such as Ca(OH)2 and NaOH, are the main post-treatment methods [5,27]. Two of the benefits of this method are the low cost and elevated stability; however, it has drawbacks such as an elevated energy demand for solvent regeneration, high corrosiveness, elevated toxicity, volatility, and high cost [20,22]. At present, the energy demand for CO2 capture at the Boundary Dam and Petra Nova projects is around 0.25–0.30 MW·h·t−1 CO2, resulting in a loss in energy efficiency. Alternatively, CO2 storage solutions based on KOH and K2CO3 have been considered [22]. Aqueous solutions of amino acids and ionic liquids have been investigated [28,29]. Compared with amine-based technology, amino acid-based technology has the advantages of low toxicity, high biodegradation potential, and low evaporation rate [30]. However, further works on the economic feasibility and degradation viability of amino acids are required [20,30]. Currently, ionic liquids have been classified as viable alternatives to the traditional physical solvents utilized in the abovementioned procedure because of their basic behaviors, including low volatility, vapor pressure, elevated thermal stability at high temperatures, and low requirements, including regeneration demand energy. Nonetheless, their low operational capacity is a major barrier to their common use in carbon capture [5,16].

2.2.2. Adsorption-Based CO2 Capture

The adsorption-based CO2 capture process includes the bulk phase transfer of CO2 molecules in the flue gas to a solid surface at atmospheric pressure and high temperature (Figure 4). The attached CO2 molecules, called adsorbates, form a thin film on the surface of a solid material, called an adsorbent. In addition, adsorption-based CO2 capture has attracted extensive interest due to its simple process, low energy consumption, and low cost. However, the low recovery of CO2 remains a challenge [5,32,33]. CO2 capture on porous solids is a promising way to extract CO2 from flue gas in different production methods, if the adsorption mechanism is based on the affinity of CO2 for adsorption on the adsorbate surface [20]. Various adsorbents have been studied for CO2 capture from pre- and post-combustion flue gases. Normally, the adsorbents adopted are categorized as elevated-temperature or low-temperature materials. The main category of elevated-temperature materials encompasses hydrotalcites, alkali metal, or alkaline earth metal oxides such as calcium oxide, alkali metal silicates and zirconates, and double salts, while low-temperature adsorbents include traditional porous materials such as zeolites and carbon-based materials [8,34,35,36,37,38]. These include activated carbon, carbon nanotubes, carbon nanofibers, graphene, and molecular sieves, in addition to newer classes such as MOFs, porous polymer networks (PPNs), and covalent organic frameworks (COFs) [5,34]. Adsorption is the most promising method, compared with absorption and other industrial methods, because it has the following advantages: (i) better performance [39], (ii) long-term stability [32], (iii) lowest regeneration cost [33], (iv) does not regenerate toxic substances [40], and (v) is inexpensive. There are two main types: (i) physisorption (van der Waals attraction) and (ii) chemisorption (interaction of chemical bonds) [32,33,39]. In the physisorption process, molecules of CO2 attach to the pore walls of the adsorbent through van der Waals and pole–pole interactions [33,41]. The heat of adsorption values for the physisorption process is close to the sublimation heat, and the values are in the range of 25–40 kJ/mol [33,42,43]. This is considered a reversible process, where the adsorption/desorption of the gas molecules can be achieved under the effect of temperature and pressure [33,44]. The CO2 adsorption at ambient temperature mainly occurs by physisorption, and the gas capture is directly related to the porous texture of the adsorbent surface [33,45,46]. It has been reported that narrow micropores (0.33–1 nm) are responsible for the CO2 adsorption performance [33,47]. To overcome the low gas selectivity for CO2 in physisorbents, chemical adsorption is widely performed on the surface of the porous materials by incorporating basic groups that effectively interact with the acidic CO2 gas molecules [33,48,49,50]. Among different basic groups, the amine is commonly used for the surface modification of CO2 adsorbents [33,41]. The CO2 gas molecule forms a chemical bond with the adsorption sites in chemisorption. Usually, CO2 adsorption at elevated temperatures (above 140 °C) is mainly ruled by chemisorption [33,45,51]. Both chemisorption and physisorption can occur in the temperature range of 25–140 °C. The heat of adsorption for chemisorbents can vary between 60 to 100 kJ/mol depending on the chemical functionality and bonding nature.

2.2.3. Membrane-Based CO2 Separation

In membrane-based gas separation, the gas flow is commonly precipitated by a pressure differential through the membrane, and separation is achieved when the flue gas permeates across the pores of the membrane (Figure 5). The process combines membrane separation and amine absorption, raising the mass transfer area (movement of mass) by 400–1500%. In addition, gas separation offers several advantages, such as energy efficiency, cost effectiveness, ease of operation, scalability, and environmental friendliness, compared with other CO2 capture methods [16,53]. Polymer-based films show promise due to their ease of processing and mechanical properties. However, they suffer from inherent problems, such as the competition between gas permeability and selectivity, in addition to poor durability (e.g., low heat and chemical resistance), plasticization, and aging, which affect their gas separation performance [20,54]. Inorganic membranes have drawbacks such as elevated cost and difficult manufacturing. Mixed-matrix membranes (MMMs) incorporate a polymer matrix with functional features such as zeolites, carbon molecular sieves (CMS), and carbon nanotubes. In terms of challenges and prospects, membranes are generally expensive due to the large quantity of energy required to pressurize the gas. Clearly, more research is needed for material development and composition design [16,20].

2.2.4. Chemical Looping-Based CO2 Capture

Chemical looping applies indirect combustion of fossil fuels and renewable fuels, dividing the combustion procedure into a two-stage process (Figure 6). First, the fuel reduces metal oxides (an oxidant that transfers oxygen into the fuel) to the metal reduction column. The resulting metals (e.g., iron, manganese, or copper) are then oxidized in an oxidation tower. Metal oxides are formed as a result of the reaction. These compounds are transported to a second reactor, where they react with the fuel. Metal oxides are reduced during combustion, generating energy and flue gas as CO2 and H2O streams. Flue gas can be condensed to obtain pure CO2 [56,57]. The system (chemical looping-based CO2 uptake) involves two interconnected reactors, named the air reactor (AR) and the fuel reactor (FR), where oxidation and reduction reactions, respectively, occur. In the FR, the gaseous fuel (CnH2mOp) reacts with the oxygen carrier (MexOy), which is reduced to a metal (Me) or a reduced form (MexOy−1), while the fuel is transformed to CO2 + H2O, according to the reaction in Equation (1). Then, the reduced oxygen carrier is transferred to the AR, where it is oxidized by oxygen, according to the reaction in Equation (2). Thus, this process occurs in two stages, avoiding direct contact between the fuel and the air [57].
2 n + m p M x O y +   C n H 2 m O p 2 n + m p M x O y 1 + nCO 2 + mH 2 Δ H r
2 n + m p M x O y 1 + n + m 2 p 2   O 2 2 n + m p M x O y Δ H 0
A key feature of this technology is the emission of exhaust gas, the main component of which is carbon dioxide, which can be recuperated by cooling. The technique has shown good results for further development, and various ceramic materials, such as NiO/NiAl2O4, have been investigated [20,56]. The large-scale application of this method is dependent on the availability of air carriers. For chemical looping operations, some essential metal oxides’ properties are important, such as high chemical activity, mechanical stability (in fluidized bed), resistance to caking, high melting point (to resist reaction temperatures and avoid caking), long-term stability under repeated oxidation/reduction, and cost [5,58].
Figure 6. CO2 capture with chemical looping combustion of gaseous fuels (adapted from [59]).
Figure 6. CO2 capture with chemical looping combustion of gaseous fuels (adapted from [59]).
Cleantechnol 04 00073 g006

3. Recent Developments and Potential Uses

CO2 has been utilized in various sectors and processes, as shown in Figure 7. Carbon dioxide is used in the chemical, oil, and power industries, in the food processing and pharmaceutical sector, and in the pulp and paper steel sector, among other uses [7,60].
The greatest global capacity for CO2 utilization is by the chemical and petroleum industries, with the greatest potential for enhanced oil recovery (EOR) (estimated up to 300 Mtpa), urea synthesis, polymer manufacturing processes for fuels, and chemical synthesis of, e.g., sustainable alcohol and acids [61,62,63]. The cement industry also has a huge absorption potential of up to 300 Mtpa. For the food industry, moderate potential exists, mainly in beverage carbonation, packaging, decaffeination, and horticulture [60,64]. If the supercritical carbon dioxide cycle in the power industry is not limited to today’s existing geothermal power plants and heat pump applications, overseas markets for potential future applications are expected to include nuclear power, coal, biomass, and natural gas to waste incinerators, such as small thermal power cogeneration application [60,65]. The CO2 utilization potential is estimated to be around 300 Mtpa [60]. The Global Carbon Dioxide Initiative provides a USD 0.8–1.1 trillion annual market for carbon-based products, as it consumes more than 10% of the world’s annual carbon dioxide emissions [60,63].
There are two other approaches to the use of carbon dioxide: direct use of carbon dioxide (physical) and conversion of carbon dioxide into useful chemicals [7,66]. However, the scope for direct use of CO2 is limited and the impact on overall CO2 reduction is also limited. The chemical conversion is carried out using CO2 as a soft oxidant [7]. The potential uses are the oxidative dehydrogenation of ethylbenzene, ethane, and propane, with CO2 used as a smooth oxidant; the transformation of CO2 to organic polycarbonates and cyclic carbonates by adding CO2 into phenol; and the transformation of CO2 to alcohols, such as methanol, and acids, such as formic acid [7,67].
In recent years, CO2 has been utilized in several industrial processes [6]. There are several technical approaches to transform CO2 into consumer products and suitable procedures, such as catalytic, photocatalytic, photosynthetic, electrochemical, mineralization, and biological (using microorganisms) processes [68]. The electrochemical process uses an electrolyzer (an apparatus that produces hydrogen through a chemical process (electrolysis) capable of separating the hydrogen and oxygen molecules) to reduce CO2 to CO. H2 is generated by the electrolysis of water and is a common reactant for converting CO2 to CH4, CH3OH, etc. The electrolysis process is expensive and requires a high energy demand [68].
Dongliang et al. [69] utilized green hydrogen to produce methanol through carbon dioxide hydrogenation; instead of releasing the greenhouse gases into the atmosphere, methanol obtained was diverted to a local project (1.8 Mt/y). In addition, Aresta [70] demonstrated that ethanol derived from atmospheric CO2 should have the same benefits as bioethanol, with the added benefit that its synthesis produces fewer pollutants than those produced by fermenting sugars. Peng et al. [71] applied photocatalysts to transform CO2 to fuel by a solar-driven process.
Chen et al. [72] studied an efficient solid CO2 adsorbent with a hierarchical micro-mesoporous structure and a large number of amine groups with high affinity for carbon dioxide, achieving CO2 capture of 3.32 mmol CO2·g−1 adsorbent at a temperature range of 303–343 K; compared with other materials used in the literature, this is a promising value. Additionally, Peres et al. [36] researched CO2 adsorption using porous carbon by sugarcane bagasse waste feedstock (the process included carbonization, followed by activation with KOH); they achieved a CO2 capture of 1.22 mmol CO2·g−1 adsorbent at room temperature. In addition, Ma et al. [35] performed the synthesis of nitrogen-doped porous carbons from polyacrylonitrile fibers as efficient CO2 adsorbents via a simple two-step synthesis process (carbonization followed by KOH activation) and achieved a CO2 removal of 3.95 mmol CO2·g−1 adsorbent at room temperature.
Another interesting study by Cho et al. [73] designed a process for recycling desalination wastewater, utilizing CO2 as CaCO3 and MgCO3 in the CCU process. The results showed that about 91% of the CO2 in the flue gas was captured and used. In addition, the maximum global profit of the proposed process was 73.23 USD per ton of CO2, which means it is profitable and suitable for large-capacity processes, compared with other alternative CCU processes. Baena-Moreno et al. [74] proposed a new process for carbon dioxide capture and utilization appropriate for small and medium-sized applications. They utilized potassium and calcium wastes as alternative low-energy pathways for carbon capture and waste recycling, promoting a circular carbon economy. Hernandéz et al. [75] reported, for the first time, a feasible process for converting CO2 to cyclic carbonate (a reagent that can be prepared using carbon dioxide), which can even improve catalytic activity and avoid the use of energy-consuming technologies by solving the problem of catalyst separation from reaction products and compatibility with reaction steps [18]. Additionally, Yao et al. [76] focused on the analysis of the technical and economic viability of addressing CO2 capture in the context of the steel sector. CO2 can be utilized in catalytic processes with CO2 reduction methods. One of these processes is the photocatalytic reduction of carbon dioxide. The photocatalytic method of CO2 photoreduction starts with light absorption in semiconductors, which allows electron-hole pairs to form in nanocrystals. These charge carriers migrate to the surface of the material and reduce or oxidize the adsorbed molecules [77]. The following reactions occur:
CO2 + e → CO2 (−1.90 V)
CO2 + 2H+ + 2e → HCOOH (−0.61 V)
CO2 + 2H+ + 2e → CO + H2O (−0.53 V)
CO2 + 4H+ + 4e → CH2O + H2O (−0.48 V)
CO2 + 6H+ + 6e → CH3OH + H2O (−0.38 V)
CO2 + 8H+ + 8e → CH4 + 2H2O (−0.24 V)
2H2O + 4H+ → O2 + 4H+ (0.81 V)
2H+ + 2e → H2 (−0.42 V)
One of the major benefits of photocatalytic processes is the direct use of photons, rather than the initial transformation to power. However, these procedures are elaborate and involve several methods, such as electron and proton transfer and bond formation [68].
Celaya et al. [77] examined the conversion of CO2 to hydrocarbons via a photocatalytic process on M-doped titanate nanotubes (formed from amorphous TiO2 using a relatively mild hydrothermal reaction conditions (0.05–1 M NaOH, 110 °C), forming parahydrophobic surfaces) (M = Ni and Cu). Ma et al. [78] proposed a feasible strategy for CO2 reduction (photocatalytic) with H2O in an aerobic condition by the high CO2/O2 adsorption selectivity of polymers, and effective charge separation for CO2 reduction and H2O oxidation on Pd (II) sites and hollow TiO2, respectively. Sun et al. [79] investigated the viability of indium oxide-supported silver catalysts in the hydrogenation of CO2 to methanol conversion by density functional theory (DFT). Alternatively, biocatalytic conversion of CO2 has received great attention. Gomes et al. [80] reported a biocatalytic cascade transformation of CO2 to methanol by enzymes co-immobilized in a tailored silicon-containing mesostructured porous foam. Additionally, Rasouli et al. [81] presented a hybrid enzymatic carbon dioxide absorption process in an enhanced flat-membrane contactor with immobilized human carbonic anhydrase II (hCA II) enzyme.
In the electrochemical field, advances have been made in CO2 capture, mainly involving quinones. Hemmatifar et al. [82] presented a feasible and customizable electrochemical method for direct CO2 capture by a bipolar cell, which utilizes the affinity of redox-active quinone moieties towards CO2 molecules.

4. Future Opportunities

In the future, CO2 technologies have great potential to limit CO2 global emissions, but there are challenges to be faced. In this context, several studies have focused on CO2 technologies and have received great attention. Gupta and Li [20] summarized several studies of CCU processes using machine learning (ML). Orlov et al. [83] produced an original computational approach combining kinetic experiments, molecular simulations, and machine learning for the screening of hundreds of potential solvent candidates and identified a class of tertiary amines that capture CO2 faster than typical solvents. In the same way, Jin et al. [84] established two different regression models based on machine learning methods to estimate carbon dioxide levels in solvents. The methods used were an adaptive augmented support vector (SVR) machine and an adaptive augmented Gaussian process (GPR) to approximate the carbon dioxide loading in solvents for environmental applications. Zhu et al. [85] designed a quantitative structure–property relationship using a machine learning model, considering 6244 CO2 adsorption datasets of 155 porous carbon materials, to predict CO2 adsorption capture and examine the importance of physicochemical effects. The results showed that the random forest (RF) model presented promising accuracy and prediction performance based on the physicochemical proprieties and adsorption conditions of porous carbon materials and the test dataset (R2 > 0.9). All these studies are significant because mathematical simulations of CO2 capture can save on time (laboratory and preparation) and costs (reagents).
Another important tool that is promising for CO2 technologies is the life cycle assessment (LCA). LCA considers the whole life cycle of products and processes (cradle to grave thinking), from raw material extraction (cradle) and transportation, through fabrication and product utilization, to recycling and final disposal of waste (grave). In this context, several life cycle assessments have been carried out for different CCU processes, which were mainly supported by assumptions [68]. He et al. [86] studied the environmental benefits of harnessing carbon dioxide by back-converting water gas to syngas to produce liquid fuels and generate electricity. The CO2 source is an ideal tertiary hydrogen synthesis unit-supported iron-based chemical loop combustion, which can absorb carbon with almost no energy demand. Life cycle emissions were measured at 129.98 kg CO2 eq.·MWh−1, corresponding to a 60% share of CO2 emissions from liquid fuel production (using exergy distribution as an example). In this system, CO2 capture using nearly zero energy could be achieved using a three-stage Fe-based chemical looping combustion (CLC) hydrogen generation system (CLHG) to obtain the CO2 source, which benefited from the advantages of CLC to realize inherent carbon capture. In the fuel reactor (FR), CH4 is indirectly combusted to generate CO2 and steam. Then, the reduced form of OC (FeO) is subsequently oxidized by steam to generate H2 in a steam reactor (SR). Finally, the partially oxidized OC (Fe3O4) from the SR is further oxidized by air to its original form (Fe2O3) in an air reactor (AR). In addition, within the framework of “cradle to cradle” thinking, Battuecas et al. [87] studied a new method to produce nano-calcium carbonate as a cement filler using carbon dioxide from cement waste gas. In this work, two pathways for CO2 capture were evaluated, followed by the synthesis of CaCO3 via mineral carbonation. Three protocols for synthesizing CaCO3 nanoparticles were evaluated, aiming to use waste as raw materials and recycle them to reduce emissions. The three scenarios were considered using a life cycle analysis approach. Rosental et al. [88] provided cradle-to-gate-based life cycle assessments to produce a wide range of organic chemicals: methanol, ethylene, propylene, benzene, toluene, and mixed xylenes. The LCA results showed that the production of the studied chemicals by the CCU process will decrease greenhouse gas emissions by 88–97% when using electricity from offshore wind turbines compared to fossil-based synthesis routes. Simultaneously, other environmental impacts, such as eutrophication processes and ozone depletion, will increase [89,90,91,92]. All these studies are significant, given that the LCA can calculate economic values, relative environmental impacts, and feasibility, and identify the most important factors.

5. Conclusions

Decreasing carbon dioxide emissions and slowing the greenhouse effect is essential for human life, global environmental safety, and sustainable development. In this short review, the state of the art and future opportunities of CCU are discussed from a holistic perspective. Certainly, recent years have seen significant progress in terms of the development of several CCU technologies. CO2 can be utilized in several processes in the chemical and oil industry (urea production, fuels), the food industry (refrigeration, food processing), mineralization (baking soda, CO2 concrete curing), power (heat pumps), energy crops (algae cultivation), pharmaceuticals (chemical synthesis), pulp and paper steel (injection to metal casting), and more (e.g., water treatment). In recent years, captured CO2 has been converted into marketable products, and CCU technology has come a long way. Many methods for the preparation of fuels/chemicals by CO2 copolymerization via catalysis (photocatalysis and biocatalysis), electrochemistry, and the biotransformation of CO2 are already present in the local market. Going forward, CCU will continue to make temporary progress, mainly in more technologically advanced areas such as CO2-derived polymers, CO2 carbonization, and methanol fabrication. In the long term, CCU will be a crucial element of a circular carbon economy. For the future, technologies such as machine learning and life cycle assessment are promising.

Author Contributions

L.C.d.M., C.B.P., P.M.R.R. and L.J.R.N.; methodology, L.C.d.M., C.B.P., P.M.R.R. and L.J.R.N.; validation, L.C.d.M., C.B.P., P.M.R.R. and L.J.R.N.; formal analysis, L.C.d.M., C.B.P., P.M.R.R. and L.J.R.N.; investigation, L.C.d.M., C.B.P., P.M.R.R. and L.J.R.N.; resources, L.C.d.M., C.B.P., P.M.R.R. and L.J.R.N.; data curation, L.C.d.M., C.B.P., P.M.R.R. and L.J.R.N.; writing—original draft preparation, L.C.d.M., C.B.P., P.M.R.R. and L.J.R.N.; writing—review and editing, L.C.d.M., C.B.P., P.M.R.R. and L.J.R.N.; visualization, L.C.d.M., C.B.P., P.M.R.R. and L.J.R.N.; supervision, L.C.d.M., C.B.P., P.M.R.R. and L.J.R.N. All authors have read and agreed to the published version of the manuscript.

Funding

L.J.R.N. was supported by Prometheus—Research Unit on Energy, Materials and Environment for Sustainability—UIDP/05975/2020, funded by national funds through FCT—Fundação para a Ciência e Tecnologia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are available upon request to the correspondent author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Peres, C.B.; Resende, P.R.; Nunes, L.J.R.; de Morais, L.C. Circular economy: A comprehensive review of eco-friendly wollastonite applications. Sustainability 2022, 14, 3070. [Google Scholar] [CrossRef]
  2. International Energy Agency. Global CO2 Emissions Rebounded to Their Highest Level in History in 2021. Available online: https://www.iea.org/news/global-co2-emissions-rebounded-to-their-highest-level-in-history-in-2021 (accessed on 31 July 2022).
  3. Perathoner, S.; van Geem, K.M.; Marin, G.B.; Centi, G. Reuse of CO2 in energy intensive process industries. Chem. Commun. 2021, 57, 10967. [Google Scholar] [CrossRef] [PubMed]
  4. Renforth, P.; Sick, V.; Sheehan, S.W.; Pace, G. Scaling CO2 capture with downstream flow CO2 conversion to ethanol. Front. Clim. 2021, 1, 656108. [Google Scholar] [CrossRef]
  5. Al-Mamoori, A.; Krishnamurthy, A.; Rownaghi, A.A.; Ezaei, F. Carbon capture and utilization update. Energy Technol. 2017, 5, 834–849. [Google Scholar] [CrossRef] [Green Version]
  6. Gao, W.; Liang, S.; Wang, R.; Jiang, Q.; Zhang, Y.; Zheng, Q.; Xie, B.; Ying Toe, C.; Zhu, X.; Wang, J.; et al. Industrial carbon dioxide capture and utilization: State of the art and future challenges. Chem. Soc. Rev. 2020, 49, 8584. [Google Scholar] [CrossRef]
  7. Challa, P.; Paleti, G.; Madduluri, V.R.; Gadamani, S.B.; Pothu, R.; Burri, D.R.; Boddula, R.; Perugopu, V.; Kamaraju, S.R.R. Trends in emission and utilization of CO2: Sustainable feedstock in the synthesis of value-added fine chemicals. Catal. Surv. Asia 2022, 26, 80–91. [Google Scholar] [CrossRef]
  8. Ding, M.; Liu, X.; Ma, P.; Yao, J. Porous materials for capture and catalytic conversion of CO2 at low concentration. Coord. Chem. Rev. 2022, 465, 214576. [Google Scholar] [CrossRef]
  9. Liu, J.; Zhang, Q.; Li, H.; Chen, S.; Teng, F. Investment decision on carbon capture and utilization (CCU) technologies—A real option model based on technology learning effect. Appl. Energy 2022, 322, 119514. [Google Scholar] [CrossRef]
  10. Wich, T.; Lueke, W.; Deerberg, G.; Oles, M. Carbon2Chem®-CCU as a step toward a circular economy. Front. Energy Res. 2020, 7, 162. [Google Scholar] [CrossRef] [Green Version]
  11. Costa, L.P.D.M.; de Miranda, D.M.V.; de Oliveira, A.C.; Falcon, L.; Pimenta, M.S.S.; Bessa, I.G.; Wouters, S.J.; Andrade, M.; Pinto, J. Capture and reuse of carbon dioxide (CO2) for a plastics circular economy: A review. Processes 2021, 9, 759. [Google Scholar] [CrossRef]
  12. Dimitriou, I.; García-Gutiérrez, P.; Elder, R.H.; Cuéllar-Franca, R.M.; Azapagic, A.; Allen, R.W. Carbon dioxide utilisation for production of transport fuels: Process and economic analysis. Energy Environ. Sci. 2015, 8, 1775–1789. [Google Scholar] [CrossRef] [Green Version]
  13. Garcia-Garcia, G.; Fernandez, M.C.; Armstrong, K.; Woolass, S.; Styring, P. Analytical review of life-cycle environmental impacts of carbon capture and utilization technologies. ChemSusChem 2021, 14, 995–1015. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, S.; Liu, J.; Zhang, Q.; Teng, F.; McLellan, B.C. A critical review on deployment planning and risk analysis of carbon capture, utilization, and storage (CCUS) toward carbon neutrality. Renew. Sustain. Energy Rev. 2022, 167, 112537. [Google Scholar] [CrossRef]
  15. Kamkeng, A.D.N.; Wang, M.; Hu, J.; Du, W.; Qian, F. Transformation technologies for CO2 utilisation: Current status, challenges and future prospects. Chem. Eng. J. 2021, 409, 128138. [Google Scholar] [CrossRef]
  16. Pahija, E.; Golshan, S.; Blais, B.; Boffito, D.C. Perspectives on the process intensification of CO2 capture and utilization. Chem. Eng. Process. Process Intensif. 2022, 176, 108958. [Google Scholar] [CrossRef]
  17. Mikulčić, H.; Skov, I.R.; Dominković, D.F.; Wan Alwi, S.R.; Manan, Z.A.; Tan, R.; Duić, N.; Mohamad, S.N.H.; Wang, X. Flexible carbon capture and utilization technologies in future energy systems and the utilization pathways of captured CO2. Renew. Sustain. Energy Rev. 2019, 114, 109338. [Google Scholar] [CrossRef]
  18. Ravikumar, D.; Zhang, D.; Keoleian, G.; Miller, S.; Sick, V.; Li, V. Carbon dioxide utilization in concrete curing or mixing might not produce a net climate benefit. Nat. Commun. 2021, 12, 1–13. [Google Scholar] [CrossRef]
  19. Zhang, S.; Liu, L.; Zhang, L.; Zhuang, Y.; Du, J. An optimization model for carbon capture utilization and storage supply chain: A case study in Northeastern China. Appl. Energy 2018, 231, 194–206. [Google Scholar] [CrossRef]
  20. Gupta, S.; Li, L. The potential of machine learning for enhancing CO2 sequestration, storage, transportation, and utilization-based processes: A brief perspective. Miner. Met. Mater. Soc. 2022, 74, 414–428. [Google Scholar] [CrossRef]
  21. Cao, L.; Lei, S.; Guan, Y.; Wang, Y.; Zhang, Y.; Tian, J.; Wang, T.; Luo, B.; Ren, T. CCUS industry under target of carbon-peak and carbon-neutrality: Progress and challenges. Front. Energy Res. 2022, 10, 860665. [Google Scholar] [CrossRef]
  22. Lin, Q.; Zhang, X.; Wang, T.; Zheng, C.; Gao, X. Technical perspective of carbon capture, utilization, and storage. Engineering 2022, 14, 27–32. [Google Scholar] [CrossRef]
  23. Wienchol, P.; Korus, A.; Szlęk, A.; Ditaranto, M. Thermogravimetric and Kinetic study of thermal degradation of various types of municipal solid waste (MSW) under N2, CO2 and oxy-fuel conditions. Energy 2022, 248, 123573. [Google Scholar] [CrossRef]
  24. Yew, S.; Chai, W.; How, B.S. Review of carbon capture absorbents for CO2 utilization. Greenh. Gases Sci. Technol. 2022, 12, 1–34. [Google Scholar] [CrossRef]
  25. Ozkan, M.; Akhavi, A.A.; Coley, W.C.; Shang, R.; Ma, Y. Progress in carbon dioxide capture materials for deep decarbonization. Chem 2022, 8, 141–173. [Google Scholar] [CrossRef]
  26. Ribeiro, M.G.; Hisse, D.; Prado, M.L.; dos Santos, T.C.; Ronconi, C.M. Sustainable technologies of CO2 capture: A brief review. Rev. Virtual De Química 2022, 14, 517–528. [Google Scholar] [CrossRef]
  27. Zhang, Y.; Wang, S.; Feng, D.; Gao, J.; Dong, L.; Zhao, Y.; Sun, S.; Huang, Y.; Qin, Y. Functional biochar synergistic solid/liquid-phase CO2 capture: A review. Energy Fuels 2022, 36, 2945–2970. [Google Scholar] [CrossRef]
  28. Aghel, B.; Janati, S.; Wongwises, S.; Shadloo, M.S. Review on CO2 capture by blended amine solutions. Int. J. Greenh. Gas Control 2022, 119, 103715. [Google Scholar] [CrossRef]
  29. Hu, G.; Smith, K.H.; Wu, Y.; Mumford, K.A.; Kentish, S.E.; Stevens, G.W. Carbon Dioxide capture by solvent absorption using amino acids: A review. Chin. J. Chem. Eng. 2018, 26, 2229–2237. [Google Scholar] [CrossRef]
  30. Zhang, Y.; Zhu, C.; Chu, C.; Fu, T.; Ma, Y. Mass transfer and capture of carbon dioxide using amino acids sodium aqueous solution in microchannel. Chem. Eng. Process. Process Intensif. 2022, 173, 108831. [Google Scholar] [CrossRef]
  31. Øi, L.; Procedia, S.K.-E. Comparison of energy consumption for different CO2 absorption configurations using different simulation tools. Energy Procedia 2014, 63, 1186–1195. [Google Scholar] [CrossRef]
  32. Akinola, T.E.; Bonilla Prado, P.L.; Wang, M. Experimental studies, molecular simulation and process modelling\simulation of adsorption-based post-combustion carbon capture for power plants: A state-of-the-art review. Appl. Energy 2022, 317, 119156. [Google Scholar] [CrossRef]
  33. Gunawardene, O.H.P.; Gunathilake, C.A.; Vikrant, K.; Amaraweera, S.M. Carbon Dioxide capture through physical and chemical adsorption using porous carbon materials: A review. Atmosphere 2022, 13, 397. [Google Scholar] [CrossRef]
  34. Farmahini, A.H.; Krishnamurthy, S.; Friedrich, D.; Brandani, S.; Sarkisov, L. Performance-Based screening of porous materials for carbon capture. Chem. Rev. 2021, 121, 10666–10741. [Google Scholar] [CrossRef] [PubMed]
  35. Ma, C.; Bai, J.; Hu, X.; Jiang, Z.; Wang, L. Nitrogen-doped porous carbons from polyacrylonitrile fiber as effective CO2 adsorbents. J. Environ. Sci. 2023, 125, 533–543. [Google Scholar] [CrossRef] [PubMed]
  36. Peres, C.B.; Rosa, A.H.; de Morais, L.C. CO2 adsorption of bagasse waste feedstock using thermogravimetric analyses. J. Therm. Anal. Calorim. 2021, 147, 5973–5984. [Google Scholar] [CrossRef]
  37. Zheng, Y.; Wu, J.; Zhang, L.; Guo, Y.; Xu, Z.; Huang, Y.; Huang, P.; Zhang, J.; Zhao, C. Unravelling the pore templating effect on CO2 adsorption performance of alkali metal nitrates promoted MgO pellets. Chem. Eng. J. 2022, 450, 137944. [Google Scholar] [CrossRef]
  38. Wang, J.; Zhou, Y.; Hu, X. Adsorption of CO2 by a novel zeolite doped amine modified ternary aerogels. Environ. Res. 2022, 214, 113855. [Google Scholar] [CrossRef]
  39. Maniarasu, R.; Rathore, S.K.; Murugan, S. Biomass-based activated carbon for CO2 adsorption—A review. Energy Environ. 2022, 0, 0958305X2210934. [Google Scholar] [CrossRef]
  40. Lu, T.; Li, Q.; Shao, J.; Wang, L.; Pang, R.; Wu, X.; Hu, X. Nitrogen and sulfur co-doped porous carbons from polyacrylonitrile fibers for CO2 adsorption. J. Taiwan Inst. Chem. Eng. 2021, 128, 148–155. [Google Scholar] [CrossRef]
  41. Cabriga, C.K.C.; Clarete, K.V.B.; Zhang, J.A.T.; Pacia, R.M.P.; Ko, Y.S.; Castro, J.C. Evaluation of biochar derived from the slow pyrolysis of rice straw as a potential adsorbent for carbon dioxide. Biomass Convers. Biorefinery 2021, 1, 3. [Google Scholar] [CrossRef]
  42. Berger, A.H.; Bhown, A.S. Comparing physisorption and chemisorption solid sorbents for use separating CO2 from flue gas using temperature swing adsorption. Energy Procedia 2011, 4, 562–567. [Google Scholar] [CrossRef] [Green Version]
  43. Wang, J.; Yuan, X.; Deng, S.; Zeng, X.; Yu, Z.; Li, S.; Li, K. Waste polyethylene terephthalate (PET) plastics-derived activated carbon for CO2 capture: A route to a closed carbon loop. Green Chem. 2020, 22, 6836–6845. [Google Scholar] [CrossRef]
  44. Nandi, M.; Uyama, H. Exceptional CO2 adsorbing materials under different conditions. Chem. Rec. 2014, 14, 1134–1148. [Google Scholar] [CrossRef] [PubMed]
  45. Gunathilake, C.; Manchanda, A.S.; Ghimire, P.; Kruk, M.; Jaroniec, M. Amine-modified silica nanotubes and nanospheres: Synthesis and CO2 sorption properties. Environ. Sci. Nano 2016, 3, 806–817. [Google Scholar] [CrossRef]
  46. Yao, M.; Wang, L.; Hu, X.; Hu, G.; Luo, M.; Fan, M. Synthesis of Nitrogen-doped carbon with three-dimensional mesostructures for CO2 capture. J. Mater. Sci. 2015, 50, 1221–1227. [Google Scholar] [CrossRef]
  47. Benedetti, V.; Cordioli, E.; Patuzzi, F.; Baratieri, M. CO2 Adsorption study on pure and chemically activated chars derived from commercial biomass gasifiers. J. CO2 Util. 2019, 33, 46–54. [Google Scholar] [CrossRef]
  48. Ochedi, F.O.; Liu, Y.; Adewuyi, Y.G. State-of-the-art review on capture of CO2 using adsorbents prepared from waste materials. Process Saf. Environ. Prot. 2020, 139, 1–25. [Google Scholar] [CrossRef]
  49. Chen, C.; Kim, J.; Ahn, W.S. CO2 Capture by amine-functionalized nanoporous materials: A review. Korean J. Chem. Eng. 2014, 31, 1919–1934. [Google Scholar] [CrossRef]
  50. Zhao, R.; Liu, L.; Zhao, L.; Deng, S.; Li, S.; Zhang, Y. A comprehensive performance evaluation of temperature swing adsorption for post-combustion carbon dioxide capture. Renew. Sustain. Energy Rev. 2019, 114, 109285. [Google Scholar] [CrossRef]
  51. Vinodh, R.; Babu, C.M.; Abidov, A.; Palanichamy, M.; Jang, H.T. Facile synthesis of amine modified silica/reduced graphene oxide composite sorbent for CO2 adsorption. Mater. Lett. 2019, 247, 44–47. [Google Scholar] [CrossRef]
  52. Sullivan, J.M.; Sivak, M. Carbon Capture in Vehicles: A Review of General Support, Available Mechanisms, and Consumer Acceptance Issues; University of Michigan, Transportation Research Institute: Ann Arbor, MI, USA, 2012. [Google Scholar]
  53. Chen, T.Y.; Deng, X.; Lin, L.C.; Ho, W.S.W. New sterically hindered polyvinylamine-containing membranes for CO2 capture from flue gas. J. Membr. Sci. 2022, 645, 120195. [Google Scholar] [CrossRef]
  54. Demir, H.; Aksu, G.O.; Gulbalkan, H.C.; Keskin, S. MOF membranes for CO2 capture: Past, present and future. Carbon Capture Sci. Technol. 2022, 2, 100026. [Google Scholar] [CrossRef]
  55. Amiri, A.; Shahbazian-Yassar, R. Recent progress of high-entropy materials for energy storage and conversion. J. Mater. Chem. 2021, 9, 782–823. [Google Scholar] [CrossRef]
  56. Gomes, P.; Madejski, P.; Chmiel, K.; Subramanian, N.; Ku’s, T.K. Methods and techniques for CO2 capture: Review of potential solutions and applications in modern energy technologies. Energies 2022, 15, 887. [Google Scholar] [CrossRef]
  57. Araújo, R.; Barros Do Nascimento, A.; Pimenta De Macedo, H.; Melo, D.M.A.; Santiago, R.C.; Rodrigues De, T.; Araújo, A.; Luiz, R.; Medeiros, B.A.; Adánez, J. Structure and Reactivity of brazilian iron ores as low-cost oxygen carriers for chemical looping combustion. Eng. Chem. Res. 2022, 61, 2469–2482. [Google Scholar] [CrossRef]
  58. Adánez-Rubio, I.; Samprón, I.; Izquierdo, M.T.; Abad, A.; Gayán, P.; Adánez, J. Coal and biomass combustion with CO2 capture by CLOU process using a magnetic Fe-Mn-supported CuO oxygen carrier. Fuel 2022, 314, 122742. [Google Scholar] [CrossRef]
  59. Li, J.; Zhang, H.; Gao, Z.; Fu, J.; Ao, W.; Dai, J. CO2 Capture with chemical looping combustion of gaseous fuels: An overview. Energy Fuels 2017, 31, 3475–3524. [Google Scholar] [CrossRef]
  60. Koytsoumpa, E.I.; Bergins, C.; Kakaras, E. The CO2 economy: Review of CO2 capture and reuse technologies. J. Supercrit. Fluids 2018, 132, 3–16. [Google Scholar] [CrossRef]
  61. Zhang, K.; Li, S.; Liu, L. Optimized foam-assisted CO2 enhanced oil recovery technology in tight oil reservoirs. Fuel 2020, 267, 117099. [Google Scholar] [CrossRef]
  62. Jiang, J.; Rui, Z.; Hazlett, R.; Lu, J. An integrated technical-economic model for evaluating CO2 Enhanced oil recovery development. Appl. Energy 2019, 247, 190–211. [Google Scholar] [CrossRef]
  63. Valluri, S.; Claremboux, V.; Kawatra, S. Opportunities and challenges in CO2 utilization. J. Environ. Sci. 2022, 113, 322–344. [Google Scholar] [CrossRef] [PubMed]
  64. Norhasyima, R.S.; Mahlia, T.M.I. Advances in CO2 utilization technology: A patent landscape review. J. CO2 Util. 2018, 26, 323–335. [Google Scholar] [CrossRef]
  65. Reiter, G.; Lindorfer, J. Evaluating CO2 sources for power-to-gas applications—A case study for Austria. J. CO2 Util. 2015, 10, 40–49. [Google Scholar] [CrossRef]
  66. Huang, C.H.; Tan, C.S. A review: CO2 utilization. Aerosol Air Qual. Res. 2014, 14, 480–499. [Google Scholar] [CrossRef] [Green Version]
  67. Centi, G.; Perathoner, S.; Salladini, A.; Iaquaniello, G. Economics of CO2 utilization: A critical analysis. Front. Energy Res. 2020, 8, 567986. [Google Scholar] [CrossRef]
  68. Zhu, Q. Developments on CO2-utilization technologies. Clean Energy 2019, 3, 85–100. [Google Scholar] [CrossRef] [Green Version]
  69. Dongliang, W.; Wenliang, M.; Huairong, Z.; Guixian, L.; Yong, Y.; Hongwei, L. Green hydrogen coupling with CO2 utilization of coal-to-methanol for high methanol productivity and low CO2 emission. Energy 2021, 231, 120970. [Google Scholar] [CrossRef]
  70. Aresta, M. Do bio-ethanol and synthetic ethanol produced from air-captured CO2 have the same degree of “Greenness” and relevance to “Fossil C”? Molecules 2022, 27, 2223. [Google Scholar] [CrossRef]
  71. Peng, W.; Nguyen, T.H.C.; Nguyen, D.L.T.; Wang, T.; Tran, T.V.T.; Le, T.H.; Le, H.K.; Grace, A.N.; Singh, P.; Raizadaa, P.; et al. A roadmap towards the development of superior photocatalysts for solar-driven CO2-to-fuels production. Renew. Sustain. Energy Rev. 2021, 148, 111298. [Google Scholar] [CrossRef]
  72. Dibenedetto, A.; Chen, Y.; Wu, J.; Wang, X.; Liu, M.; Liu, Y. Synthesis, characterization and application of amine-functionalized hierarchically micro-mesoporous silicon composites for CO2 capture in flue gas. Molecules 2022, 27, 3429. [Google Scholar] [CrossRef]
  73. Cho, S.; Lim, J.; Cho, H.; Yoo, Y.; Kang, D.; Kim, J. Novel process design of desalination wastewater recovery for CO2 and SOX utilization. Chem. Eng. J. 2022, 433, 133602. [Google Scholar] [CrossRef]
  74. Baena-Moreno, F.M.; Vega, F.; Pastor-Pérez, L.; Reina, T.R.; Navarrete, B.; Zhang, Z. Novel Process for carbon capture and utilization and saline wastes valorization. J. Nat. Gas Sci. Eng. 2020, 73, 103071. [Google Scholar] [CrossRef]
  75. Hernández, E.; Santiago, R.; Moya, C.; Vela, S.; Navarro, P.; Palomar, J. Close-cycle process to produce CO2-derived propylene carbonate based on amino acid catalyst and water. J. CO2 Util. 2021, 52, 101656. [Google Scholar] [CrossRef]
  76. Yao, X.; Yuan, X.; Yu, S.; Lei, M. Economic feasibility analysis of carbon capture technology in steelworks based on system dynamics. J. Clean Prod. 2021, 322, 129046. [Google Scholar] [CrossRef]
  77. Celaya, C.A.; Méndez-Galván, M.; Castro-Ocampo, O.; Torres-Martínez, L.M.; Luévano-Hipólito, E.; de León, J.N.D.; Lara-García, H.A.; Díaz, G.; Muñiz, J. Exploring the CO2 conversion into hydrocarbons via a Photocatalytic process onto M-Doped Titanate Nanotubes (M = Ni and Cu). Fuel 2022, 324, 124440. [Google Scholar] [CrossRef]
  78. Ma, Y.; Yi, X.; Wang, S.; Li, T.; Tan, B.; Chen, C.; Majima, T.; Waclawik, E.R.; Zhu, H.; Wang, J. Selective photocatalytic CO2 reduction in aerobic environment by microporous Pd-Porphyrin-based polymers coated hollow TiO2. Nat. Commun. 2022, 13, 1–10. [Google Scholar] [CrossRef]
  79. Sun, K.; Zhang, Z.; Shen, C.; Rui, N.; Liu, C.-J. The Feasibility study of the indium oxide supported silver catalyst for selective hydrogenation of CO2 to methanol. Green Energy Environ. 2021, 7, 807–817. [Google Scholar] [CrossRef]
  80. Gomes, M.Z.D.V.; Masdeu, G.; Eiring, P.; Kuhlemann, A.; Sauer, M.; Åkerman, B.; Palmqvist, A.E.C. Improved biocatalytic cascade conversion of CO2 to methanol by enzymes co-immobilized in tailored siliceous mesostructured cellular foams. Catal. Sci. Technol. 2021, 11, 6952. [Google Scholar] [CrossRef]
  81. Rasouli, H.; Iliuta, I.; Bougie, F.; Garnier, A.; Iliuta, M.C. Hybrid enzymatic CO2 capture process in intensified flat sheet membrane contactors with immobilized carbonic anhydrase. Sep. Purif. Technol. 2022, 287, 120505. [Google Scholar] [CrossRef]
  82. Hemmatifar, A.; Kang, J.S.; Ozbek, N.; Tan, K.-J.; Hatton, T.A. Electrochemically mediated direct CO2 capture by a stackable bipolar cell. ChemSusChem 2022, 22, e202102533. [Google Scholar] [CrossRef]
  83. Orlov, A.A.; Valtz, A.; Coquelet, C.; Rozanska, X.; Wimmer, E.; Marcou, G.; Horvath, D.; Poulain, B.; Varnek, A.; de Meyer, F. Computational screening methodology identifieseffective solvents for CO2 capture. Commun. Chem. 2022, 5, 1–7. [Google Scholar] [CrossRef]
  84. Jin, H.; Andalib, V.; Yasin, G.; Bokov, D.O.; Kamal, M.; Alashwal, M.; Ghazali, S.; Algarni, M.; Mamdouh, A. Computational simulation using machine learning models in prediction of CO2 absorption in environmental applications. J. Mol. Liq. 2022, 358, 119159. [Google Scholar] [CrossRef]
  85. Zhu, X.; Tsang, D.C.W.; Wang, L.; Su, Z.; Hou, D.; Li, L.; Shang, J. Machine learning exploration of the critical factors for CO2 adsorption capacity on porous carbon materials at different pressures. J. Clean. Prod. 2020, 273, 122915. [Google Scholar] [CrossRef]
  86. He, Y.; Zhu, L.; Fan, J.; Li, L.; Liu, G. Life cycle assessment of CO2 emission reduction potential of carbon capture and utilization for liquid fuel and power cogeneration. Fuel Process. Technol. 2021, 221, 106924. [Google Scholar] [CrossRef]
  87. Batuecas, E.; Liendo, F.; Tommasi, T.; Bensaid, S.; Deorsola, F.A.; Fino, D. Recycling CO2 from flue gas for CaCO3 nanoparticles production as cement filler: A life cycle assessment. J. CO2 Util. 2021, 45, 101446. [Google Scholar] [CrossRef]
  88. Rosental, M.; Fröhlich, T.; Liebich, A. Life cycle assessment of carbon capture and utilization for the production of large volume organic chemicals. Front. Clim. 2020, 2, 586199. [Google Scholar] [CrossRef]
  89. Khoo, Z.Y.; Ho, E.H.Z.; Li, Y.; Yeo, Z.; Low, J.S.C.; Bu, J.; Chia, L.S.O. Life cycle assessment of a CO2 mineralisation technology for carbon capture and utilisation in Singapore. J. CO2 Util. 2021, 44, 101378. [Google Scholar] [CrossRef]
  90. Yaïci, W.; Entchev, E.; Longo, M. Recent Advances in Small-Scale Carbon Capture Systems for Micro-Combined Heat and Power Applications. Energies 2022, 15, 2938. [Google Scholar] [CrossRef]
  91. Arena, N.; Lee, J.; Clift, R. Life cycle assessment of activated carbon production from coconut shells. J. Clean. Prod. 2016, 125, 68–77. [Google Scholar] [CrossRef] [Green Version]
  92. Gu, H.; Bergman, R.; Anderson, N.; Alanya-Rosenbaum, S. Life cycle assessment of activated carbon from woody biomass. Wood Fiber Sci. 2018, 50, 229–243. [Google Scholar] [CrossRef]
Figure 1. Scheme of CCU technologies (adapted from [6,20]).
Figure 1. Scheme of CCU technologies (adapted from [6,20]).
Cleantechnol 04 00073 g001
Figure 2. CO2 capture and separation technology (adapted from [16,20]).
Figure 2. CO2 capture and separation technology (adapted from [16,20]).
Cleantechnol 04 00073 g002
Figure 3. Principle for a standard CO2 removal process based on absorption followed by desorption in amine solution (adapted from [31]).
Figure 3. Principle for a standard CO2 removal process based on absorption followed by desorption in amine solution (adapted from [31]).
Cleantechnol 04 00073 g003
Figure 4. Carbon dioxide separation by adsorption (adapted from [52]).
Figure 4. Carbon dioxide separation by adsorption (adapted from [52]).
Cleantechnol 04 00073 g004
Figure 5. Membrane separation technology in carbon capture (adapted from [55]).
Figure 5. Membrane separation technology in carbon capture (adapted from [55]).
Cleantechnol 04 00073 g005
Figure 7. CO2 utilization by sector (adapted from [60]).
Figure 7. CO2 utilization by sector (adapted from [60]).
Cleantechnol 04 00073 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Peres, C.B.; Resende, P.M.R.; Nunes, L.J.R.; Morais, L.C.d. Advances in Carbon Capture and Use (CCU) Technologies: A Comprehensive Review and CO2 Mitigation Potential Analysis. Clean Technol. 2022, 4, 1193-1207. https://doi.org/10.3390/cleantechnol4040073

AMA Style

Peres CB, Resende PMR, Nunes LJR, Morais LCd. Advances in Carbon Capture and Use (CCU) Technologies: A Comprehensive Review and CO2 Mitigation Potential Analysis. Clean Technologies. 2022; 4(4):1193-1207. https://doi.org/10.3390/cleantechnol4040073

Chicago/Turabian Style

Peres, Christiano B., Pedro M. R. Resende, Leonel J. R. Nunes, and Leandro C. de Morais. 2022. "Advances in Carbon Capture and Use (CCU) Technologies: A Comprehensive Review and CO2 Mitigation Potential Analysis" Clean Technologies 4, no. 4: 1193-1207. https://doi.org/10.3390/cleantechnol4040073

Article Metrics

Back to TopTop