Next Article in Journal
Embodied Environmental and Social Impacts: A Regionalised Sectoral Method for Low-Carbon Construction Materials in Italy
Previous Article in Journal
The Impacts of Green Supply Chain Management and Product Innovation on Marketing Performance in Thailand’s Processed Food Industry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Low-Energy Regeneration Technologies for Industrial CO2 Capture: Advances, Challenges, and Engineering Applications

1
School of Energy and Environmental Engineering, University of Science and Technology Beijing, Beijing 100083, China
2
Capital Engineering & Research Incorporation Limited, Beijing 100176, China
3
Beijing SDL Technology Co., Ltd., Beijing 102206, China
*
Authors to whom correspondence should be addressed.
Sustainability 2025, 17(21), 9796; https://doi.org/10.3390/su17219796
Submission received: 3 September 2025 / Revised: 13 October 2025 / Accepted: 30 October 2025 / Published: 3 November 2025

Abstract

High carbon dioxide (CO2) emissions from industrial processes have intensified the need for large-scale, sustainable, and low-energy-consumption carbon capture technologies. Amine-based chemical absorption is a promising method for large-scale CO2 reduction, but it faces challenges like high regeneration energy consumption, technical limitations, and commercialization difficulties. To reduce energy consumption in regeneration, this paper reviews low-energy regeneration methods, including absorbent optimization, catalytic regeneration, process waste heat recovery, and calcium-based chemical desorption, and explains the energy-saving mechanisms of each approach. Focusing on technical development bottlenecks, this paper provides a comprehensive review of the technical advantages, application limitations, and key challenges associated with various methods. Based on commercialization needs, this paper thoroughly investigates the development process and industrialization status of carbon capture technology in the iron and steel industry. Research has revealed that optimized absorbent designs reduce regeneration heat loads, catalytic acid sites promote proton transfer and lower desorption temperature, utilization of waste heat reduce additional energy consumption, and calcium-based compounds offer both low energy consumption and economic advantages in desorption. This article constructs a theoretical framework for low-energy regeneration technology, identifies innovation priorities, and analyzes scalability challenges and development pathways, providing theoretical support and technical guidance for industrial implementation.

1. Introduction

The rising concentration of greenhouse gases in the atmosphere is one of the key factors contributing to global warming and climate change. Over the past few decades, rapid industrialization and urbanization have brought unprecedented economic development and population growth. It has led to a dramatic consumption of fossil fuels, such as coal, oil, and natural gas, which are the primary sources of greenhouse gases, particularly CO2 [1]. Currently, the concentration of CO2 in the atmosphere has exceeded 425 ppm and continues to rise [2]. 2024 has become the warmest year since global records began in 1850. To successfully limit global temperature rise to 1.5 °C above pre-industrial levels, CO2 emissions must be significantly reduced, particularly in the industrial sector [3,4]. Carbon Capture, Utilization, and Storage (CCUS) is considered one of the key technologies for mitigating climate change by capturing CO2 [5], from large point sources, and subsequently reusing it, such as converting CO2 into high-value-added chemicals through catalysis [6,7], like methanol [8,9], etc., or storing it.
Carbon capture technologies mainly include: pre-combustion capture, mid-combustion capture, and post-combustion capture technologies [10]. Pre-combustion capture involves separating CO2 from fossil fuels before combustion starts, converting the fuel into a gas mixture of H2, CO, H2O, and CO2 through gasification, thereby achieving CO2 separation and removal [11]. However, this process requires high energy consumption and costs, limiting its widespread application. Oxy-fuel combustion is the most widely used mid-combustion capture technology, which puts the fuel in a high concentration of oxygen or pure oxygen combustion rather than direct using air, to make the fuel fully combusted, the main combustion products are CO2 and H2O [12]. Therefore, it can increase the CO2 concentration in the flue gas after combustion, facilitating subsequent CO2 capture. Oxygen-enriched combustion can be implemented through simple modifications to traditional combustion equipment, with low investment costs. However, this technology requires a large amount of O2 during operation, resulting in high operating costs. Additionally, the presence of water vapor can cause corrosion to the equipment [12,13]. Post-combustion carbon capture technology, which does not interfere with or alter existing process flows, offers high adaptation and is widely applied in industrial sectors [14].
Post-combustion carbon capture technology uses absorption, adsorption, and membrane separation technologies to directly capture CO2 from flue gas that has pre-treated with denitrification, desulfurization, and dust removal [15]. Captured CO2 typically requires subsequent regeneration prior to utilization or storage. Chemical absorption, distinguished by its high reaction rates and substantial CO2 absorption capacity, is one of the most prevalent and mature technologies for CO2 capture, and has been used for decades for CO2 removal [4,16]. In the chemical absorption process, CO2 in flue gas reacts with absorbents (e.g., amines, carbonates) to form intermediate compounds, after which the solvent is regenerated through heating or other methods. Amine-based absorbents demonstrate high efficacy in capturing CO2 from flue gases and other emission streams, where stable chemical compounds are formed following CO2 absorption [17]. Monoethanolamin (MEA) is regarded as the benchmark absorbent due to its elevated reaction kinetics and low cost [18]. Nevertheless, its regeneration process requires heating to above 100 °C, which consumes a large amount of energy and increases the cost of technical application to a certain extent. Therefore, it is necessary to further reduce regeneration energy consumption through methods such as synthesizing new absorbents [17].
In 2023, global carbon emissions from the industrial sector will be approximately 9207 Mt, with the steel industry contributing over 30% and ranking first [19]. In 2019, nearly 74.5% of crude steel production came from coal-based steel plants, mainly in China [20]. Although dynamic modeling analysis has found that production restructuring can promote carbon neutrality in the industry [21], production restructuring poses certain challenges in the short term due to the maturity and widespread use of the blast furnace-basic oxygen furnace (BF-BOF) process [22,23]. Therefore, CCUS plays a key role in carbon emission reduction in the steel industry [24]. In 2016, ADNOC Al-Reyadah Company completed and operated the first carbon capture project in the steel industry in the United Arab Emirates, with an annual CO2 capture capacity of 800,000 tons [25]. It uses the MEA amine process, whereby captured CO2 is compressed and transported to enhanced oil recovery sites, and ultimately stored underground [26]. However, regenerating absorbents requires high energy consumption, which increases the cost of carbon capture for enterprises. Although many scholars have conducted research on low-energy-consumption regenerated absorbents, their studies have been limited to system simulation or laboratory analysis [27].
Although amine chemical absorption technology is widely applied in CO2 capture, its high energy consumption during regeneration poses a significant bottleneck that severely limits its economic viability. Furthermore, existing research lacks a systematic review of low-energy regeneration technologies. To address this, this paper first comprehensively reviews research progress on low-energy regeneration technologies. It systematically outlines the energy-saving mechanisms, application effectiveness, advantages, limitations, and challenges of 4 primary technical pathways—absorbent optimization, catalytic regeneration, technology integration, and chemical desorption—establishing a theoretical framework for low-energy regeneration in CO2 chemical absorption. Addressing the immense emission reduction pressure and limited application of carbon capture projects in the steel industry—a typical hard-to-abate sector—this paper further analyzes the current development status and commercialization progress of carbon capture technologies in this industry. In order to provide reference and insights for carbon capture practices in other heavy industrial sectors. On this basis, it looks at the future of low-energy recycling tech, figuring out the challenges and ways to achieve large-scale application. This review establishes a systematic theoretical framework for low-energy CO2 chemical absorption regeneration technology, aiming to provide important theoretical support and technical guidance for the innovative development and engineering application of the technology. The specific research framework of this paper is shown in Figure 1.

2. Overview of CO2 Capture Technology

Industrial sources (such as the steel industry) emit large amounts of CO2 into the environment during production operations. Therefore, carbon capture technology is needed to reduce and control CO2 emissions [16]. To reduce the impact of CO2 emissions, CCUS technology is regarded as the best solution to the problem [13,16,19]. Among them, post-combustion capture is considered the best option among existing capture technologies [13]. Based on the principles of the capture process, the main capture technologies include absorption, adsorption, membrane separation, and other novel technologies. Among these, absorption, adsorption, and membrane separation have received widespread attention in this field. In addition, emerging carbon capture technologies such as low-temperature technology, calcium cycling, and chemical looping combustion have also attracted extensive research.

2.1. Absorption

Absorption is a relatively mature technology that has already been commercialized. It mainly includes physical absorption and chemical absorption. Physical absorption is based on Henry’s law, which absorbs CO2 under high pressure and low temperature conditions, and regenerates the solvent under low pressure or high temperature conditions, or a combination of both [28]. Physical absorption operates based on the solubility differences between the solvent and CO2 as well as other gases. There are several physical absorption methods in industry, such as Rectisol, Selexol, Purisol, and Fluor [29]. Common physical solvents include methanol, dimethyl ether of polyethylene glycol, N-methyl-2-pyrrolidone, and propylene carbonate [13,19,30]. Physical absorption requires low-temperature and high-pressure conditions. In industrial environments, interference from impurity gases, low capture rate of low-concentration CO2, and high solvent regeneration energy consumption limit its widespread application [30,31]. Chemical absorption involves using a weak alkaline solution to react chemically with CO2 to form weakly bound compounds, which are then regenerated through heating or other means. Due to its simple process flow, high absorption efficiency, high maturity, and low cost, it is one of the most widely used technologies [32]. The main chemical absorbents include: amine solvents, carbonates, hydrated ammonia, and amino acids, etc. [13,14,19]. Among these, amine-based absorbents exhibit a unique affinity for CO2, forming stable compounds after reaction, which can be reused after thermal regeneration. They are considered the most practical method for controlling CO2 emissions from industrial flue gas combustion [15,32].
Common amino compounds include: MEA, diethylethanolamine (DEEA), and N-Methyldiethanolamine (MDEA) [33]. The most classic amine absorbent is 20–30 wt% water-based MEA [15,16,19]. The general process for CO2 absorption is shown in Figure 2. After combustion, flue gas containing CO2 contact with a lean amine solution in an absorption tower, where the amine solution selectively reacts with CO2 to form a CO2-rich amine solution. This solution is then sent to a stripper for solvent regeneration and CO2 desorption, resulting in high-concentration CO2 for further utilization or storage.
MEA is a primary amine, and its specific reaction process with CO2 is as follows [14,35]. First, amines react with CO2 to form zwitterion, which has a faster reaction rate (Equation (1)). After that, the amine molecule reacts with zwitterion as a Bronsted base (proton acceptor) to produce protonated amines and carbamates (Equation (2)). At high pressure, carbamates may undergo hydrolysis, yielding free amine and bicarbonate (Equation (3)). According to the reaction process, the theoretical CO2 loading was 0.5 mol CO2/mol amine.
R 1 R 2 N H + C O 2 ( a q ) R 1 R 2 NH + C O O
R 1 R 2 N H + R 1 R 2 NH + C O O R 1 R 2 NH + + R 1 R 2 N C O O
R 1 R 2 N C O O + H 2 O R 1 R 2 N H + H C O 3
The reaction mechanism of secondary amines is the same as that of primary amines. However, tertiary amines cannot react directly with CO2 because there are no hydrogen atoms on the nitrogen atom. But they can provide proton acceptors to promote the hydrolysis of CO2 to form bicarbonate (Equation (4)) [36]. Compared with primary and secondary amines, their theoretical CO2 loading capacity is 1 mol CO2/mol amine, but their reactivity is lower.
R 1 R 2 R 3 N + C O 2 ( a q ) + H 2 O R 1 R 2 R 3 N H + + H C O 3
Due to its high reaction rate, high CO2 absorption capacity, and low cost, MEA is considered the benchmark absorbent [18]. However, the regeneration of CO2-rich MEA solutions requires a significant amount of energy, which increases the cost of CO2 capture and hinders the industrial application of chemical absorption. Scholars have already mixed other solvents into MEA to enhance its CO2 absorption capacity and reduce regeneration energy consumption. By mixing a small amount of tertiary amine (MDEA) into MEA to form a solvent mixture, they fully utilize the advantages of both, thereby improving the solvent’s capture performance and reducing the regeneration energy demand [35].
Chemical absorption using amine solvents is currently the most mature technology for CO2 capture and has been commercially applied for many years. It offers advantages such as high reaction rates, high CO2 loading capacity, and high recovery rates. However, during operation, the solvent undergoes degradation and evaporation, leading to solvent loss and equipment corrosion issues. Additionally, since the compounds formed after capturing CO2 with amine-based solvents have high stability, regeneration requires temperatures exceeding 100 °C, resulting in significant energy consumption during the regeneration process. To reduce regeneration energy consumption, further development of new, highly efficient absorbents is necessary.

2.2. Adsorption

The adsorption method for capturing CO2 involves gas being adhered to a solid porous surface, thereby being captured by the adsorbent. Solid adsorption serves as an alternative to carbon capture technologies, offering advantages such as reduced energy consumption, high adsorption capacity, good selectivity, and simple operation compared to amine-based chemical absorption methods [13,19]. Similarly to absorption, adsorption is also divided into physical adsorption and chemical adsorption [13]. Physical adsorption involves the adsorption of CO2 through physical adsorbents under intermolecular forces. Common physical adsorbents include activated carbon, zeolites, and carbon nanotubes, etc. [37,38]. based on their main chemical components. Since no new chemical bonds are formed, the energy consumed in regenerating the adsorbent is lower than that in chemical adsorption, but this also results in poorer selectivity in physical adsorption [19]. Zeolite has higher selectivity than other materials, but its CO2 loading capacity is low, and its efficiency decreases in the presence of water vapor [38]. Metal–organic frameworks (MOFs) have certain advantages in carbon capture due to their functionalized pore morphology and customized structures. After modification with the introduction of active sites, MOFs exhibit high CO2 loading capacities [39]. However, desorption requires high temperatures or low pressures, which increases regeneration energy consumption.
Chemical adsorption has stronger selectivity than physical adsorption, mainly because modified solid materials can form new chemical bonds with CO2. The main types of modified materials include two categories: organic amines [40] and inorganic metal oxides (alkali metals or alkaline earth metals) [37]. Most chemical adsorbents are prepared by impregnation, which allows active sites to be uniformly distributed on the adsorbent [41]. Jialiang Li et al. [40] analyzed the mechanism of interaction between amine-functionalized graphene oxide and CO2 through simulation, and found that there exists physical adsorption between the two via van der Waals forces and hydrogen bonds, as well as chemical adsorption between the amine groups and CO2. Shaoliang Zhu et al. [42] co-impregnated two amine solutions on γ-Al2O3 and found that under the synergistic action of the two amines, the maximum CO2 loading was 2.65 mmol/g. The CO2 adsorption process mainly includes three steps: external diffusion, internal diffusion, and adsorption reaction. CO2 enters the interior of the adsorbent through external diffusion, then reaches the active adsorption sites via internal diffusion, subsequently undergoing a physico-chemical reaction. The CO2 adsorption reaction process of amine-functionalized adsorbents is illustrated in Figure 3.
Under anhydrous conditions, the main reaction processes are shown in Equations (5) and (6) [44]:
R 1 R 2 NH + C O 2 R 1 R 2 N H + C O O
R 1 R 2 N H + C O O + R 1 R 2 NH R 1 R 2 NH 2 + + R 1 R 2 N C O O
In the presence of water, the reaction process is similar to chemical absorption, and the reaction rate is faster than under anhydrous conditions [43].
In the CO2 capture process, organic amine adsorbents exhibit high selectivity and are suitable for low-concentration CO2 capture [45]. Additionally, they have advantages such as lower regeneration energy consumption compared to water-based amine absorbents. However, in practical applications, the adsorbent may become deactivated or degraded under the influence of CO2, O2, acidic gases, and heating, thereby affecting its CO2 capture efficiency [43]. Furthermore, their relatively low CO2 adsorption capacity and high cost limit their industrial-scale application.

2.3. Membrane Separation Technology

Membrane separation technology is a process that uses semipermeable membranes to separate mixed gases composed of two or more components [46]. During membrane separation, the membrane acts as a filter, allowing specific molecules (such as CO2) to pass through, thereby enriching CO2 on the other side of the membrane. CO2 permeates through the membrane via two mechanisms: facilitated transport and solution-diffusion, while other gases only pass through the membrane via the solution-diffusion mechanism, as shown in Figure 4. Its ability to separate CO2 depends on the permeability and selectivity of the membrane material to CO2 [47]. Depending on the composition of the material, common membrane materials used for CO2/N2 separation after combustion include polymers, ceramics, zeolites, and MOFs [13].
Polymer materials are widely used due to their heat resistance, chemical resistance, excellent film-forming ability, and processability [13,49]. The flue gas produced after combustion mainly contains gases such as CO2, H2O, and N2. The separation of CO2 and N2 mainly relies on surface diffusion and solution diffusion, which is driven by differences in adsorption capacity and solubility between the gases. Since the diffusion selectivity of CO2/N2 in polymers is similar, CO2/N2 separation membranes enhance their separation capabilities by introducing functional groups that can undergo reversible reactions with CO2 [50]. For selective membranes containing functional groups, CO2 dissolves in the polymer on the high-pressure side and reacts with the functional groups. The reaction products diffuse along the concentration gradient on both sides of the membrane. On the low-pressure side, the reaction products undergo a reversible dissociation reaction, releasing CO2. Related studies have found that functional groups containing nitrogen, oxygen, sulfur, or phosphorus can increase the affinity between CO2 and the membrane [51]. Although there are already a large number of polymer membrane materials available, polymer membrane materials exhibit trade-off effects in gas separation processes, making it difficult to achieve both permeability and selectivity simultaneously [49]. In addition, polymer membranes face issues such as physical aging and plasticization induced by CO2, leading to a decline in separation performance and selectivity [49].
To overcome the limitations of polymers, researchers prepared hybrid matrix membranes by adding inorganic materials to polymers [52]. It fully leverages the low cost of organic membranes and the high permeability, selectivity, and stability of inorganic materials. Currently, there are various hybrid matrix membranes synthesized using different inorganic materials, such as carbon nanotubes, carbon molecular sieves, zeolite molecular sieves, and MOFs [47]. Jiangnan Wang et al. [53] used the MOF material Zeolitic Imidazolate Framework-8 (ZIF-8) to synthesize a mixed matrix membrane and found that it has excellent CO2 separation performance and good selectivity. Related studies have also found that enhancing the interaction between the polymer matrix and the filler helps alleviate physical aging and plasticization phenomena [47]. Although mixed matrix membranes can improve CO2 separation performance and selectivity, problems such as agglomeration, sedimentation, and poor dispersion of fillers during the preparation process limit their further application [52].
Membrane-based CO2 capture technology offers advantages such as low cost, ease of operation, low energy consumption, and modularity [46,48]. Compared with traditional amine chemical absorption technology, it does not cause corrosion or require the handling of harmful solvents, thereby reducing the environmental impact caused by solvents and their by-products. Polymers are widely used as membrane materials, but their application in separation processes is constrained by trade-off effects. Hybrid matrix membranes can effectively improve CO2 separation performance and selectivity, but their preparation process still poses challenges. Therefore, it is necessary to further explore advanced gas separation membrane materials to prepare hybrid matrix membranes with stable performance and high selectivity.

2.4. Emerging Carbon Capture Technologies

(a)
Cryogenic technology
Cryogenic technology is a physical capture method that separates CO2 gas from flue gas by utilizing its different condensation and sublimation characteristics compared to other flue gas components. Depending on the type of phase change, it can be divided into liquefaction capture and sublimation frost capture [54]. The cryogenic process is shown in Figure 5. The flue gas is continuously cooled, causing CO2 to separate from the flue gas in liquid or solid form, while the nitrogen-rich flue gas is emitted into the atmosphere or used for cooling water [55]. Since this method employs low-temperature condensation separation, it does not introduce additional chemical reagents, thereby avoiding the byproduct pollution and equipment corrosion issues associated with chemical absorption processes [56]. Additionally, this technology achieves high CO2 purity (99.99%) and recovery rates (99.99%), making it a carbon capture technology with significant potential and practical value [13,56,57].
However, the application of low-temperature technology in the CO2 capture process presents certain challenges. Industrial flue gas typically contains gases such as SO2, NOx, and H2O, which can interfere with the CO2 cooling process and lead to equipment corrosion and blockages [60]. In addition, cryogenic technology is limited in its widespread application due to issues such as incomplete specific phase change mechanisms and the high economic costs associated with creating cryogenic conditions [19,54,56,57].
(b)
Calcium looping
The calcium cycle process captures CO2 using CaO at high temperatures, as shown in Figure 6. Due to the abundance, low cost, and environmental friendliness of its raw materials (such as limestone and dolomite) and its synergistic effects with other industrial processes (such as the cement industry), the calcium cycle process has gained widespread recognition in the field of CCUS technology [61,62,63]. The process consists of two circulating fluidized beds, named the carbonizer and the calciner [64]. CaO circulates between the two reactors. In the carbonation reactor, CaO reacts with CO2 in the flue gas to form CaCO3. The CaCO3 is then calcined in the calciner at a higher temperature to regenerate CaO and produce high-concentration CO2. The regenerated CaO recycles back into the carbonation reactor, while the concentrated CO2 can be stored and utilized. The carbonization reaction occurs at high temperatures and releases heat, enabling effective heat recovery and reducing the energy and economic costs of the carbon capture process [61].
The calcium looping process has advantages of using environmentally friendly materials, low energy consumption, and synergistic effects with other industrial processes that require CaO. But as the number of cycles increases, CaO gradually sinter, reducing the effective surface area available for reacting with CO2 and decreasing the gas–solid reaction rate, thereby lowering the CO2 capture efficiency [64]. Additionally, collisions between particles and the reactor cause CaO wear, which further affects the reaction rate. These issues slow down its industrial application [62].
(c)
Chemical looping combustion
Chemical looping combustion (CLC) is a promising CO2 capture technology that can achieve nearly 100% capture efficiency while ensuring low energy consumption [66]. The CLC system consists of two reactors: a fuel reactor (FR) and an air reactor (AR), as shown in Figure 7. Air in the system does not come into direct contact with fuel, but rather oxygen is transferred between two reactors [67]. In the air reactor, oxygen in the air reacts chemically with the oxygen carrier, causing it to oxidize. At the same time, nitrogen, residual oxygen, and other inert gases are expelled from the reactor. The oxidized oxygen carrier then enters the fuel reactor, where it undergoes a reduction reaction with the fuel, converting the oxygen carrier into metal or a lower oxidation state [19]. During this reaction, the oxygen carrier releases oxygen and reacts with the fuel, producing water and CO2 as the main reaction products [68]. Through further combustion with pure oxygen, the products of incomplete combustion are further reduced. After compression and condensation, almost pure CO2 can be obtained [66].
It avoids mixing CO2 from fuel combustion with nitrogen in the air, cutting down on the high energy costs of capturing gases later [69,70]. Additionally, since the fuel does not burn directly in the air within the reactor and the oxygen carrier is reoxidized at a lower temperature in the air, NOx production is reduced [60]. there are still some challenges to the widespread application of this technology, particularly because the oxygen carrier circulates continuously within the system, leading to a decrease in oxygen carrier loading and reaction rates [71].

2.5. Summary

The substantial CO2 emissions generated during industrial production processes require control through carbon capture technologies. The large-scale industrial application of current carbon capture technologies is constrained by technical maturity and the costs associated with construction and operation. Absorption technology is highly mature and commercially deployed, offering high absorption rates and capacities. However, its regeneration energy consumption is relatively high and requires further reduction to lower operating costs. Adsorption technology features low regeneration energy consumption and no corrosion issues. Yet, its CO2 selectivity and gas–liquid mass transfer adsorption rates are lower than those of absorption technology, and high-efficiency adsorbents are also costly. Membrane separation technology operates with low energy consumption and requires no chemical additives. However, its technological maturity is insufficient, and stability is poor. Cryogenic technology can produce high-purity CO2, but it requires cooling flue gas to extremely low temperatures, resulting in high operating costs. Calcium looping and CLC technologies have low operating costs and require no additional energy input, but they suffer from equipment wear and tear, and their CO2 capture efficiency needs improvement. In summary, absorption technology currently holds a significant advantage, but its regeneration energy consumption must be further reduced to lower operating costs.
The advantages and disadvantages of different carbon capture technologies are shown in Table 1.

3. Low-Energy Consumption Recycling Technology Research

Currently, the most widely used industrial capture technology is amine-based chemical absorption, which is mainly due to its advantages of fast reaction speed and large absorption capacity. In addition, it can be integrated with existing production equipment, making it easy for existing companies to adopt based on carbon emission reduction needs [30]. However, in practical application, the regeneration process of amine solvents consumes a large amount of energy. For the 30 wt% MEA benchmark solvent, its regeneration energy consumption reaches 3.80 GJ/t CO2, accounting for 57% to 70% of total energy demand, which to a certain extent increases the energy consumption and CO2 treatment costs of enterprises [73,74]. Therefore, many scholars have conducted research on reducing the energy consumption of CO2 regeneration in the chemical absorption process [35,74,75].

3.1. Absorbent Optimization

Absorbent optimization can significantly reduce regeneration energy consumption, thereby reducing regeneration costs. The energy consumption for amine solvent regeneration mainly includes the desorption heat of reaction products (urea, bicarbonate, carbonate), the sensible heat of heating the rich amine solution to the desorption temperature, and the vaporization heat of the solvent [74]. The regeneration process of traditional water-based amine absorbents (such as water-based MEA) not only causes problems such as amine degradation and equipment corrosion, but also results in more than 70% of the water vaporizing into steam, thereby increasing the energy consumption of the regeneration process [35,76]. Therefore, in recent related studies, scholars have optimized the development of new absorbents or mixed different amine solvents to reduce the energy consumption of the desorption process.
Tert-amines have been widely studied due to their high theoretical CO2 loading capacity and low regeneration energy consumption. Sudkanueng Singto et al. [77] synthesized 5 new tertiary amines and found that at 363 K, the regeneration rate of 2 mol/L 4-(dibutylamino)-2-butanol (DBAB) was more than twice that of MDEA, and the regeneration energy consumption was 29.8% lower than that of MDEA. While tertiary amines lack N-H bonds and cannot react directly with CO2, their aqueous solutions promote CO2 hydrolysis while acting as proton acceptors to form protonated amines and generate carbonates. It is not the same as primary or secondary amines, and the reaction heat is also lower, which means that the energy consumption for regeneration during desorption is also lower [78]. Although the synthesis of new tertiary amines can significantly reduce regeneration energy consumption, their reaction rate with CO2 is relatively slow, making them unsuitable for use alone in post-combustion CO2 capture [79]. Researchers have therefore proposed investigating the possibility of using mixed amines [35,80].
Mixed amine solutions combine the advantages of different types of amines, can simultaneously meet various requirements for CO2 absorption and desorption, and have better capture and regeneration effects than single amines, which are gradually being widely used [35,80]. By mixing 2-amino-2-methyl-1-propanol (AMP) with 2- (dimethylamino)ethanol (DMAE) in a 1:1 ratio, it exhibited high CO2 solubility (0.56 mol CO2/mol amine) and low regeneration energy consumption (at 363 K, regeneration heat load of 53.81 kJ/mol CO2) [81]. The regeneration energy consumption is 72.4% lower than that of MEA at the same concentration. Since AMP and DMAE are steric amines, they mainly generate bicarbonates after reacting with CO2, resulting in low regeneration heat load [82,83]. In addition to binary mixtures, Chikezie Nwaoha et al. [84] mixed AMP, piperazine (PZ), and MEA to form a ternary solvent, which had 42.9% lower regeneration energy consumption than MEA at 363K. Besides saving energy during regeneration, the absorption rate of the mixed amine solution system is also improved. Related research shows that the absorption rate of the mixed amine solution system is higher than the sum of the absorption rates of the tertiary amine and primary amine systems alone. The reason is that tertiary amines compete with protonated primary amines for protons, allowing more primary amines to react with CO2 [85].
Poor water/non-water-based absorbents are widely studied because they can reduce the high energy consumption caused by the high specific heat capacity of water-based amine absorbents during regeneration [86,87]. Non-aqueous absorbents are composed of amines and organic solvents, and have low corrosivity, low degradability, and low regeneration energy consumption [88]. Replacing water with organic solvents can direct the products of the reaction between CO2 and amines toward unstable carbonate compounds, thereby reducing the temperature and energy consumption of the regeneration process [87,89]. Hui Guo et al. [90] replaced water with glycol ethers when mixing with MEA, finding that this significantly reduced thermal load, approximately 55% lower than MEA aqueous solutions. Lv Bihong et al. [88] mixed AMP with 2-(2-aminoethylamino) ethanol (2-(2-aminoethylamino)ethanol, AEEA) and N-methyl pyrrolidone (NMP) to form a non-aqueous absorbent. At 393.15 K, the regeneration thermal load was 91.96 kJ/mol CO2, which was only half that of a 30 wt% water-based MEA absorbent. Related studies have analyzed that, during the initial stage of the reaction, the primary product is urethane. As the CO2 loading increases, under the presence of small amounts of water or ethanol, part of the urethane undergoes hydrolysis to form carbonates [88,91]. This phenomenon enables the water-poor/non-aqueous-based absorbent to not only reduce the vaporization heat of the aqueous-based absorbent but also lower the desorption heat of the reaction products, thereby reducing the energy consumption of the system during the regenerative process of the absorbent.
Ionic liquids (ILs) are liquid at room temperature and exhibit low volatility, high thermal stability, high CO2 solubility, and easily adjustable physicochemical properties, making them promising green solvents for CO2 capture [92,93]. They are composed of organic cations and anions, and their physical and chemical properties can be altered by modifying the ions or functional groups in the solvent. It can then be designed to enhance its CO2 absorption rate, capacity, and reduce CO2 regeneration energy consumption [92,94]. Traditional ILs can selectively capture CO2 through physical absorption in industrial gas separation processes [93]. However, their CO2 absorption capacity is not suitable for industrial flue gas CO2 capture at low pressures [95]. Task-specific ionic liquids (TSIL) are based on traditional ILs with specific functional groups (such as amine groups) introduced to enhance CO2 absorption capacity [96]. Yingying Zhang et al. [97] screened 76 ionic liquids based on theoretical CO2 absorption capacity and regeneration heat load. Among them, 1-hexyl-3-methylpyridinium bis(trifluoromethylsulfonyl)imide ([Hmpy][NTf2]) exhibited the lowest regeneration energy consumption at 323 K, with a value of 44.604 kJ/mol CO2, which is approximately 24% of that of a 30 wt% MEA solution. In addition to introducing other functional groups, ILs can also be mixed with water, amines, or other compounds to improve their CO2 capture capacity [93]. Lingdi Cao et al. [94] mixed 1-butyl-3-methyl-imidazolium nitrate ([Bmim][NO3]), MEA, and water in a mass ratio of 1:3:6 and found that although the addition of ILs slightly reduced the CO2 loading, the regeneration energy consumption of the mixture was reduced. Yang Lei et al. [98] customized a 1-methylpyridine trifluoroacetic acid ([C1Py][CF3COO])-ethanol mixed solvent and simulated the flue gas of a 600-megawatt coal-fired power plant. They found that the mixed solvent achieved a significant energy saving effect of 72.7% compared to MEA. TSIL’s synthesis and purification involve multiple steps, and its increased viscosity and high production costs limit its industrial application [96,99].
Developing new amine absorbents or mixed amine solvents can significantly reduce regeneration energy consumption. Water-based amine absorbents require high vaporization heat during regeneration, so water-poor/non-aqueous absorbents have been developed to further reduce regeneration energy consumption. In addition, ILs have been extensively studied due to their high CO2 solubility, low volatility, and easily adjustable physical and chemical properties. However, while reducing regeneration energy consumption, the viscosity of the absorbent often increases, leading to a decrease in absorption rate. Therefore, further research is needed on the corrosion resistance, absorption rate, and economic feasibility of new absorbents.

3.2. Adding Catalysts or Other Materials

Catalytic regeneration of CO2 amine solutions can improve desorption performance, lower desorption temperatures, and accelerate desorption rates, thereby reducing the thermal load of CO2 regeneration [100]. Currently, commonly used catalysts include metal oxides [101], zeolite molecular sieves [100,102], and MOF-based catalysts [103,104]. Yanni Guo et al. [101] prepared aluminum-modified SAPO-34 composite catalysts using the impregnation method, and reduced the energy consumption of CO2-rich amine solution regeneration by loading different mass ratios of Al2O3. It was found that at 361 K, the catalyst performed best when loaded with 15 wt% Al2O3, with a 78.4% increase in CO2 desorption rate and a 37% reduction in regeneration heat load. Zhan Tan et al. [105] synthesized two catalysts (1/2-SnO2/ATP and 1/4-SnO2/ATP) using the precipitation-impregnation method, with attapulgite (ATP) as the carrier and weight ratios of SnCl2·6H2O to ATP of 1:2 and 1:4, respectively. At 361 K, the desorption rate of 1/2-SnO2/ATP increased by 265%, and the regeneration heat load decreased by approximately 52%. Through mechanistic studies, it was found that the enhancement of MEA regeneration depends on Brønsted acid sites (BAS) and Lewis acid sites (LAS) on the catalyst. The catalyst provides acid sites that transfer protons to the urethane group, while the conjugate base of the acid site acts as a proton acceptor to accelerate the deprotonation process of protonated MEA, thereby reducing the high energy consumption associated with proton transfer [106].
Zeolite molecular sieves are often used as catalysts or carriers due to their large specific surface area and good thermal stability [102]. Gwan Hong Min et al. [100] used mesoporous silica SBA-15 as a carrier and loaded it with NiO to study its desorption heat load in a CO2-rich MEA solution at 86 °C. They found that SBA-15 loaded with 10 wt% NiO could reduce the desorption heat load by 19.9%. The composite catalyst SO42−/ZrO2/SBA-15, also using SBA-15 as the carrier, reduced regeneration energy consumption by up to 26.5% compared to the reference aqueous MEA solution at 370.15 K [107]. Xitong Yang et al. [108] synthesized MnOOH/HZSM-5 catalyst by loading MnOOH onto HZSM-5. It takes full advantage of the acidity of HZSM-5 and the synergistic ability of MnOOH for proton transfer, thereby promoting the process of protonated amine regeneration. When the catalyst concentration was 0.2 wt% in the amine solution, the regeneration thermal load was reduced by 31.8%.
MOFs have a high specific surface area, adjustable pore size, and diverse structures, allowing specific functional groups to be incorporated to meet functional requirements, thereby improving CO2 capture and regeneration capabilities [39]. It is a porous material composed of metal ions or metal clusters and organic ligands. Applying it to the desorption process has been found to increase the CO2 desorption rate [103]. ZIF-8 and Zeolitic Imidazolate Framework-67 (ZIF-67) materials formed by the coordination of metal ions with dimethylimidazole are widely used as amine solution regeneration carrier materials due to their high thermal stability, adjustable pore size, and broader functionality [104]. Muhammad Waseem et al. [109] loaded MoO3 onto ZIF-67 and further treated it with phosphotungstic acid to regulate the acidic sites of the catalyst. Under conditions where only 0.1 wt% of the catalyst was added as MEA, the regeneration heat load was reduced by 32% compared to the reference solution. Characterization analysis revealed that the activation by phosphotungstic acid further enriched the LAS and BAS sites on the MoO3 surface. The synergistic effect of these sites significantly reduced the thermal energy required for urethane decomposition and accelerated the CO2 desorption rate.
Catalysts primarily accelerate proton transfer processes by providing acidic sites, which reduces regeneration energy consumption and accelerates desorption rates. Since amine solutions are primarily alkaline, this results in a lack of proton participation during the decomposition of urethanes. The addition of a catalyst provides more acidic sites for CO2-rich amine solutions, accelerating the deprotonation of protonated amines and thereby reducing the high energy consumption required for the proton transfer process. However, catalysts face issues such as insufficient surface area, reduced activity, and agglomeration during operation, necessitating further exploration.

3.3. Technology Enhancement

Besides just lowering the heat load, a lot of research is looking at how to modify process equipment to cut down on regeneration energy use or overall carbon capture costs. Traditional tower-based capture systems have been studied a lot, and there has been a bunch of research on changing the reaction conditions in the system or adding different equipment for absorption and desorption [110]. In the context of process optimization, many studies utilize process simulators to simulate and determine optimal operating conditions [111]. Se-Young Oh et al. [110] simulated traditional amine-based CO2 capture configurations and 4 different types of modified structures using the UniSim® process simulator. They found that flue gas separation alone could reduce total energy consumption by 7.4% compared to the baseline process. Zhiwei Zhang et al. [112] optimized the scrubbing columns of the absorption and desorption towers, which could further reduce thermal load and capture costs. New process equipment has also been proposed to optimize the high energy consumption of the regeneration process. To address the high energy consumption required for solvent regeneration at the reboiler during desorption, Jaeheum Jung et al. [111] proposed a new desorption configuration, namely a combination of rich vapor recompression (RVR) and cold solvent split (CSS). RVR is a newly developed device to minimize preheating energy losses caused by cold solvent splitting. Cold solvent segmentation eliminates the reflux ratio in the aspirator. This combination improves RVR’s ability to evaporate cold solvents in heat exchangers, reducing the heat demand of the reboiler by 20%. Won-Hee Lee et al. [113] developed a process for CO2 capture utilizing an absorption tower and subsequent absorbent regeneration via reverse electrodialysis (RED). This system achieves energy and absorbent recovery during the desorption stage through RED. However, for primary and secondary amines, carbamates react with H2O as intermediate products through the ion exchange membrane of RED, and are replaced by amines and protonated amines, resulting in absorbent loss. Because tertiary amines do not produce charged intermediate products, they will not cause the loss of absorbents during the regeneration process, and the regeneration efficiency can reach 99.96%, which is more suitable for the application of this system. Therefore, selecting the optimal absorbent for the system can significantly save energy and have a good regeneration effect.
Waste heat utilization and new types of heat pumps are being researched to reduce the energy consumption of CO2 capture systems. Te Tu et al. [114] designed a novel CO2 regeneration process that combines gas-assisted desorption with waste heat recovery from the desorption gas. it enhances the driving force for CO2 regeneration by introducing auxiliary gas into the bottom or side of the desorption tower and utilizing the waste heat from the desorption gas (primarily water vapor). Compared to traditional thermal regeneration processes, the new regeneration process can reduce the thermal load of the regenerator by 7.9%. However, since auxiliary gas must be introduced into the system, there is still the issue of separating it from the CO2 gas. Therefore, it still requires some time before this process can be applied in practice. Ting Lei et al. [115] developed a novel heat exchanger to reduce efficiency losses. The cooling heat from flue gas, the steam at the top of the desorber, the condensate from the reboiler, and the cooling heat from compressed CO2 are recovered to reheat the heat exchanger, which is then used for solvent regeneration. Compared to traditional carbon capture processes, the energy consumption for regeneration was reduced by 18.03%.
In terms of reducing regenerative energy consumption, researchers have achieved certain results through the renovation of traditional process equipment, the development of new equipment, and the utilization of waste heat and new heat pump applications, such as reducing total energy consumption and regenerator heat load; nevertheless, some new processes face practical application challenges such as gas separation, and further research and optimization are needed in the future to achieve the widespread application of efficient and economical carbon capture technology.

3.4. Chemical Desorption

Chemical desorption, also known as CO2 mineralization, uses inexpensive calcium-based compounds (such as CaO [116], CaCl2 [117], and Ca(OH)2 [118]) to react with HCO3 in alcoholamine-rich liquid at room temperature to form carbonate precipitates. Protonated amines react with alkalis in calcium-based compounds to complete regeneration. Besides calcium-based compounds, some researchers have also studied magnesium-based compounds [119,120]. It simulates the natural weathering process of rocks, where CO2 dissolves in the atmosphere to form carbonic acid, which reacts with alkaline minerals to produce carbonates, and can permanently and safely capture and store CO2 [116,121]. Long Ji et al. [116] developed an integrated CO2 absorption-mineralization process in which the amine absorbent is regenerated chemically rather than by conventional thermal regeneration, significantly reducing regeneration energy consumption. Furthermore, the resulting CaCO3 is a non-toxic and stable compound that can be widely used as a filler in paper, rubber, and paint [122].
Chemical desorption can significantly reduce regeneration energy consumption and do not require further treatment of CO2. Weifeng Zhang et al. [118] studied the desorption rate and regeneration energy consumption of 5 CO2-rich solutions, including MEA and MDEA, by adding Ca(OH)2. Under optimal reaction conditions, the MEA solution exhibited the highest desorption rate of 85.31%, and the MEA-poor solution still had excellent absorption capacity after multiple cycles. Compared with thermal desorption, chemical desorption can be carried out at a lower temperature (40 °C), and the energy consumption for regeneration is only 21.28% of the thermal desorption. Long Ji et al. [116] added CaO to 5 common CO2-rich amine solutions and found that PZ exhibited the highest regeneration efficiency of 91%. Compared with traditional thermal desorption, chemical desorption does not require equipment for further treatment of regenerated CO2, such as desorbers, heat exchangers, and CO2 compressors, which significantly reduces equipment costs and operating costs.
In order to further reduce costs, many scholars have used industrial waste materials such as incineration fly ash and coal ash to chemically desorb CO2-rich amine solutions [116,123,124]. Long Ji et al. [116] added coal ash to a CO2-rich PZ solution and found that PZ undergoes regeneration after the reaction, but its regeneration efficiency is lower than that of CaO solution with the same calcium dosage. It is mainly due to the high heterogeneity of coal ash, which results in a lower theoretical CO2 loading capacity. Yan Wang et al. [125] used semi-dry desulfurization slag to mineralize 4 typical CO2-rich amine solutions, with AMP showing the highest desorption efficiency of 98%. Furthermore, chemical desorption using semi-dry desulfurization slag resulted in lower desorption energy consumption and CO2 capture costs than thermal desorption. Although the use of industrial waste for CO2 mineralization helps reduce CO2 emissions and promote the comprehensive utilization of industrial waste, its large-scale application still faces issues such as low mineralization efficiency and insufficient mass and heat transfer [124].
Compared with thermal desorption, chemical desorption can directly sequester CO2 and significantly reduce regeneration energy consumption, equipment and operating costs. The CaCO3 generated by the reaction also has certain commercial value. In addition, in order to further reduce the cost of desorbents, industrial waste can be used for CO2 regeneration. However, its large-scale application still faces problems such as low mineralization efficiency and insufficient mass and heat transfer. In the future, further in-depth research on its reaction kinetics is needed.

3.5. Summary

Traditional amine-based absorbents suffer from high regeneration energy consumption and high volatility issues. Therefore, researchers have analyzed CO2 regeneration energy consumption and sought to reduce it by lowering desorption heat or vaporization heat. This paper summarizes the methods used in recent years to reduce regeneration energy consumption, including adsorbent optimization, adding catalysts or other materials, technology enhancement, and chemical desorption. The advantages and disadvantages of different methods are summarized in Table 2. The high energy consumption during regeneration primarily stems from the regeneration of the CO2-rich absorbent—a 30 wt% MEA solution (benchmark solvent). Therefore, researchers first focused on optimizing the absorbent. This optimization is achieved by simultaneously or individually reducing the desorption heat and vaporization heat, which can significantly lower regeneration energy consumption. However, other issues associated with changing the absorbent must also be considered, including the high viscosity of saturated solvents, reduced absorption rates, and concerns regarding corrosion and thermal stability. Rich-amine regeneration is also a chemical reaction process; it can be optimized by adding catalysts or other materials to reduce regeneration energy consumption and temperature. Common issues with catalysts, such as insufficient surface area, deactivation, and agglomeration, require further resolution. Technology enhancement involves maximizing utilization of internal waste heat to reduce additional energy demands on the regeneration system. This requires modifying existing processes to optimize waste heat utilization. Consequently, it is more suitable for facilities not yet equipped with carbon capture units, where waste heat utilization modifications can be incorporated during design to save construction costs. Chemical desorption significantly reduces regeneration energy consumption while saving equipment costs and operational expenses. However, it necessitates continuous calcium source replenishment. Additionally, the disposal of generated CaCO3 after regeneration—including storage and transportation—requires thorough consideration. In summary, by optimizing regeneration energy consumption, CO2 regeneration energy consumption can be significantly reduced. However, in practical applications, factors such as actual flue gas conditions, regeneration temperature, and regeneration costs must be fully considered.

4. Iron and Steel Industry Carbon Capture Development

Iron and steel production heavily relies on fossil fuels to generate heat and act as a reducing agent to convert iron ore into iron, making it one of the world’s largest industrial sources of CO2 emissions [22]. Over 70% of steel production relies on the BF-BOF process. Given that steel production facilities typically have a long service life, existing plants are unlikely to be replaced by new technologies, and achieving industry-wide emissions reduction targets through measures such as improving production systems alone is not feasible [24]. Therefore, implementing carbon capture technology is the only viable method for achieving large-scale carbon emissions reductions in the steel industry. This section analyzes the current state of research and commercial application of carbon capture technologies in the steel industry.
The feasibility of applying different combustion carbon capture technologies in the steel industry has been studied. Navid Khallaghi et al. [128] used a model to simulate the technical feasibility of pre-combustion carbon capture using MDEA, finding that the CO2 capture rate from blast furnace gas ranged from 27% to 65%. Jorge Perpiñán et al. [23] considered using oxygen-enriched air as hot air for the combustion of coke and auxiliary fuels, with part of the exhaust gas recirculated back into the furnace. However, this may affect the reduction of iron oxides, and the high costs of modifying the blast furnace and pipelines limit the application of oxy-fuel combustion technology. Both pre-combustion and oxy-fuel combustion require modifications to existing production equipment or pipelines, which may disrupt normal production operations. In contrast, post-combustion carbon capture only requires minimal modifications to existing facilities, making it widely studied and applied in the steel industry [129].
Researchers are conducting studies on the application of post-combustion capture technologies such as absorption, adsorption, and membrane separation in the steel industry to verify the feasibility of these technologies. Yuhang Yang et al. [129] conducted a study on carbon capture using chemical absorption methods at a long-process (BF-BOF) steel plant with an annual crude steel production capacity of 9.7 Mt and CO2 emissions of 21 Mt/a. Using Aspen Plus v12® software for simulation, they investigated CO2 capture using MEA absorbents for blast furnace gas (BFG), coke oven gas (COG), Linz-Donawitz process gas (LDG), and lime kiln flue gas (LKG) using MEA absorbent, achieving a maximum annual CO2 reduction of 7.65 Mt. Mengxiang Fang et al. [130] studied the carbon capture capacity of MDEA and 2-(piperazin-1-yl) ethanamine (AEP) in a laboratory-scale study. They found that the mixed absorbent not only had a higher absorption rate but also had 28.5% lower regeneration energy consumption than MEA. Ying Xie et al. [27] also conducted pilot-scale testing of carbon capture from simulated blast furnace gas using a mixture of MDEA and AEP. After more than 200 h of stable operation, the energy consumption for regenerating the mixed solvent was only 2.47 GJ/t CO2, which is 34.66% lower than that of MEA solution. This demonstrates the feasibility of chemical absorption for carbon capture in the steel industry. Seokwon Yun et al. [131] conducted an economic assessment of the application of absorption and membrane separation technologies in the steel industry through modeling and simulation using Unisim® and MATLAB®. An increase in CO2 concentration reduces capture costs, with membrane separation technology showing a greater reduction. Therefore, at higher CO2 concentrations, membrane separation is more cost-effective and energy-efficient than absorption. Calin-Cristian Cormos [132] evaluated calcium looping and absorption technology for decarbonization in the steel industry and found that calcium looping technology has significant advantages in terms of CO2 emission reduction capacity and investment costs. In addition, calcium looping has the advantage of requiring fewer modifications to the steel manufacturing process and auxiliary facilities. For solid wastes such as steel slag, Kailun Chen et al. [123] proposed a leaching-mineralization-regeneration process to regenerate solvents with low energy consumption while comprehensively utilizing solid wastes and producing CaCO3, as shown in Figure 8. However, it still requires consideration of multiple factors before achieving true commercial application, as laboratory-scale and demonstration-scale operations often exhibit differences in flue gas composition, reaction temperature, degradation products, and operating conditions [133].
Commercial carbon capture projects in the steel industry are still in developing stages. In 1972, Occidental Terrell in the United States launched its first carbon capture project, which was applied to the natural gas/liquid natural gas industry [19]. However, it was not until 2016 that the first carbon capture project in the steel industry using the MEA amine process and with an annual CO2 capture capacity of 0.8 Mt was completed and put into operation [25]. In 2019, CO2 capture technology for lime kiln exhaust gas developed by Chinese steel companies was put into use, and the project has been operating smoothly. Related research has found that using pressure swing adsorption after oxy-fuel combustion in converters results in lower carbon emissions than chemical absorption [134]. In 2025, China Baotou Iron and Steel’s CCUS demonstration project successfully achieved full process integration. Using steel slag (SS) and CO2 as raw materials, the project produces high-purity negative-carbon calcium carbonate and sintering flux products, effectively utilizing approximately 30,000 t of CO2 annually. SS is a major byproduct of steel production, characterized by high production volumes and low utilization rates. Its primary components include CaO, SiO2, Fe2O3, and MgO, among others [124]. Shu-Yuan Pan et al. [135] conducted a comprehensive assessment of the potential of mineralization technology to reduce CO2 emissions, and the results showed that SS mineralization could reduce global CO2 emissions by 134.85 Mt. The use of steel slag for CO2 mineralization capture not only solves the problem of steel slag disposal, but also solves the problem of CO2 emissions and reduces the costs of CO2 capture, compression, and transportation [136]. Aside from the above-mentioned carbon capture projects currently in operation in the steel industry, there are 5 other projects under construction or evaluation globally [137]. Table 3 shows the current commercial carbon capture application projects in the steel industry.
In considering the current commercial application of CCUS technology in the steel industry, the cost of carbon capture is a critical factor. Its costs primarily stem from initial construction investment, capture expenses, and transportation and storage fees. Research indicates that a carbon capture project with an annual capture capacity of 0.1 Mt requires an investment of approximately USD 2.7 billion [24]. Operational costs for carbon capture are influenced not only by the capture technology itself but also by capture efficiency and the CO2 concentration in flue gas. Table 4 summarizes the costs of various carbon capture technologies for steel industry flue gas. The table demonstrates that flue gas CO2 concentration significantly impacts capture costs. Given the relatively low CO2 concentration emitted by the steel industry, capture costs are higher compared to those with higher CO2 concentrations. Therefore, the steel industry should select low-cost capture technologies based on its actual CO2 emission concentration when implementing carbon capture. In the future, as the energy consumption for regenerating chemical absorption technologies continues to decrease, further reductions in carbon capture costs are anticipated.
In summary, the steel industry faces a severe emissions reduction challenge as a high-carbon emitting point source, and carbon capture technology is a feasible method for large-scale emissions reduction. Among various combustion carbon capture technologies, post-combustion carbon capture has been widely studied due to its minimal requirements for retrofitting existing facilities. Post-combustion capture technologies such as absorption, adsorption, and membrane separation have demonstrated feasibility in laboratory and pilot-scale studies, but they remain some distance from commercial application. Currently, commercial carbon capture projects in the steel industry are in the development phase, with some projects already operational, such as the Baotou Steel CCUS Demonstration Project in China, which incorporates steel slag mineralization technology that combines SS disposal with CO2 emissions reduction. Further efforts are needed to promote the industrialization of these technologies to achieve deep decarbonization in the steel industry.

5. Future Directions and Challenges

5.1. Direction of Technological Innovation

In recent years, the development of high-efficiency, low-regeneration energy consumption absorbents has become a hot topic in the field of carbon capture technology research. In order to further improve absorption rates and absorption capacity while reducing regeneration energy consumption, current approaches primarily involve modifying the composition of the absorbent or developing phase-change absorbents that only regenerate the rich-phase solvent, thereby reducing reaction heat and regeneration heat [138]. Zhen Wang et al. [139] developed a low-viscosity non-aqueous absorbent and introduced MDEA as a regulator and dimethyl sulfoxide (DMSO) as a co-solvent. The study found that the lowest viscosity of the non-aqueous absorbent after CO2 absorption was 13.12 mPa·s, and the energy consumption for regeneration was also significantly reduced, 53% lower than that of MEA. Yujing Zhang et al. [140] introduced ethylene glycol (EG) into a polyamine/organic ether system prone to precipitation, developing a novel non-aqueous homogeneous absorbent based on polyamine/composite solvents. The study found that the absorbent reduced the sensible and latent heat during desorption, with regeneration energy consumption 42.8% lower than that of MEA. Non-aqueous absorbents can achieve relatively low regeneration energy consumption to a certain extent, but the increase in viscosity after absorbing CO2 leads to a decrease in reaction rate. Therefore, it is necessary to develop low-cost, low-viscosity non-aqueous absorbents that can be regenerated with low energy consumption.
Biphasic absorbents have attracted widespread attention due to their ability to significantly reduce regeneration energy consumption [141]. Yanlong Hu et al. [142] developed a novel phase change absorbent by adding n-butanol to MEA aqueous solution. At 120 °C, the regeneration energy consumption of the novel absorbent was approximately 35% lower than that of MEA aqueous solution, and the desorption volume was reduced by approximately 50%. Rui-Qi Jia et al. [143] used the ionic liquid 1-butyl-3-methylimidazolium tetrafluoroborate ([Bmim][BF4]) as a promoter, adding it to AEEA and DEEA aqueous solutions to improve phase separation ability and absorption/desorption performance. The study found that the addition of [Bmim][BF4] not only enhanced the solvent’s CO2 absorption performance but also reduced regeneration energy consumption by approximately 44% compared to MEA aqueous solutions, achieved by lowering specific heat capacity and reducing water evaporation. Xinling Zhong et al. [144] synthesized a novel non-aqueous ionic liquid biphasic absorbent, which has a regeneration energy consumption approximately 51% lower than that of MEA aqueous solutions. Yao Shen et al. [141] analyzed the relationship between the properties and performance of amines in two-phase solvents and thermal load. They found that the alkalinity of tertiary amines determines the absorption capacity of two-phase solvents, while the hydrophobicity of primary amines primarily affects phase separation behavior. Therefore, two-phase solvents can be tailored by adding appropriate amines to achieve high absorption capacity, good phase separation performance, and low regeneration energy consumption. Sihong Cheng et al. [126] conducted a comprehensive review of the research progress on phase change absorbents. The high viscosity of phase change absorbents requires high regeneration energy consumption during the regeneration process, thereby increasing the capture cost. Thus, choosing a low-viscosity solvent is conducive to further reducing regeneration energy consumption.
Currently, carbon capture technology focuses on developing high-efficiency, low-regeneration energy consumption absorbents. Both non-aqueous absorbents and two-phase absorbents can significantly reduce regeneration energy consumption, but the viscosity increases after absorbing CO2, which affects the reaction rate and also causes an increase in regeneration energy consumption. Further exploration of new absorbents or additives can be conducted to reduce the viscosity of the rich liquid and promote the transition of experimental research to industrial applications.

5.2. Challenges of Large-Scale Application

Although there are currently a variety of carbon capture technologies available, they are all based on laboratory-scale or system simulations. Emerging technologies need to undergo a series of scaling steps before they can be commercialized, and therefore face many challenges [16]. Primarily, the carbon capture process requires a certain amount of energy consumption, which imposes additional energy and material demands on factories. It increases operating costs for enterprises and also results in additional carbon emissions [26]. Carbon emissions avoided is commonly used as an effective indicator of carbon emission reduction to measure the difference in total carbon emissions released into the atmosphere [26]. Taking the steel industry as an example, the cost of carbon capture using MEA aqueous solutions is 52.6–73.5 USD/t CO2 [131]. Solomon Aforkoghene Aromada et al. [145] analyzed the costs of CO2 capture and CO2 avoidance in the cement industry, with CO2 capture costs ranging from 62.1 to 69.3 USD/t CO2, while the costs of avoided CO2 ranged from 74.3 to 88.9 USD/t CO2. Therefore, the implementation of carbon capture can actually reduce carbon emission costs for businesses. However, chemical absorption using organic amine solvents can result in energy consumption as high as 3.7–4.0 GJ/t CO2 during the regeneration process, leading to an increase in overall carbon capture costs and, consequently, higher production costs for the product [126].
Large-scale application of existing low-energy technologies is currently the main issue. In laboratory research, various types of absorbents have been studied. However, there has been little research on the industrial feasibility of these technologies, and most pilot-scale studies have focused on amine-based absorbents [146]. Yuli Artanto et al. [147] conducted a pilot-scale comparative study of a mixture of AMP and PZ solutions with MEA solutions. They found that when capturing higher concentrations of CO2, the energy consumption for regenerating the two absorbent liquids was similar. They further found that increasing the concentration of the absorbent increased the reaction rate with CO2, but the increase in concentration was limited by viscosity or solubility. Markus Rabensteiner et al. [148] investigated the suitability of glycine hydrate as a CO2 absorption solvent. Using real flue gas from a coal-fired power plant, they analyzed the carbon capture performance of sodium glycine hydrate at different concentrations under industrial conditions. They found that the regeneration energy consumption of 25 wt% sodium glycine hydrate was higher than that of MEA aqueous solution, with energy demand increasing by approximately 40%. In addition to assessing the overall benefits of carbon capture units, various characteristics should also be considered, including corrosiveness, degradation rate, by-products generated during the degradation process, environmental impact, and manufacturing costs [149]. Yi Ye et al. [146] developed a new type of high-load absorbent and conducted pilot-scale research. Through continuous optimization of the absorbent’s performance, the minimum regeneration energy consumption was 2.85 GJ/t CO2. Meanwhile, the low liquid-to-gas ratio demonstrated excellent potential for industrial application. However, it did not analyze the corrosion of the absorbent or the emission of by-products, so it is still some distance away from actual industrial application.
Carbon capture technology still faces challenges in terms of large-scale application. Future research needs to focus on solving the problem of large-scale application of low-energy consumption technology and thoroughly investigate the feasibility of laboratory results on an industrial scale in order to promote the practical industrial application of low-energy consumption carbon capture technology.

5.3. Policy-Driven

Policy intervention can accelerate the implementation of low-energy-consumption regenerative carbon capture projects in a variety of ways. CCUS technology is crucial to achieving carbon neutrality goals. Due to the enormous costs associated with industrial carbon capture, large-scale CCUS projects have progressed slowly. Therefore, it is necessary to drive progress through the implementation of appropriate policies. To encourage high-emission industries to adopt carbon capture technology, various policy incentives have been proposed and implemented globally. These mainly include: financial support, carbon pricing mechanisms (carbon taxes, carbon trading, and cap-and-trade systems), as well as regulatory and standard measures [150]. Countries and regions such as the United States and the United Kingdom have implemented a series of policy incentives, including direct financial support, carbon emission reduction tax credits, and carbon tax reverse incentives, to encourage industrial carbon capture projects [150]. For example, Section 45Q of the U.S. Inflation Reduction Act provides a tax credit of 85 USD/t of permanently stored carbon dioxide, offering financial incentives for businesses to capture and store carbon dioxide emissions [151]. China has also established a “1 + N” policy framework for carbon peaking and carbon neutrality, but CCUS-related policies remain primarily focused on encouragement and guidance, with no fiscal or tax incentives yet in place to support the scaled-up development of the industry. Additionally, current legal and policy frameworks are not conducive to the further development of CCUS technology [152]. Due to the current lack of CCUS demonstration projects in the steel industry, there is a gap in relevant technical standards, which has hindered technical exchanges between industries. Hence, it is necessary to further improve the legal and policy framework and expand government financial support to accelerate the widespread implementation of CCUS technology.
In the future, the global community should strengthen communication and cooperation in the formulation of CCUS policies, learn from advanced experiences, and establish a more comprehensive and targeted policy framework. On the one hand, investment in the research and development of low-energy carbon capture technologies should be increased to reduce the cost of applying these technologies. On the other hand, unified technical standards and specifications should be established as soon as possible to promote the widespread application of CCUS technologies.

6. Conclusions

Driven by both global sustainable development strategies and carbon neutrality goals, CCUS has become a core approach for reducing industrial carbon emissions and mitigating climate change. This study provides a comprehensive review of carbon dioxide capture technologies. Among these, chemical absorption technology based on amines has been widely researched and adopted due to its advantages of simple process flow, high absorption efficiency, high maturity, and high adaptability. Nevertheless, high regeneration energy consumption increases the cost of carbon capture, limiting its further commercial application. This study analyzes the main advances in current low-energy regeneration technologies. New absorbents have been synthesized to reduce regeneration energy consumption, including novel amines, mixed amine solutions, water-poor/non-aqueous-based absorbents, and ionic liquids. These primarily reduce the thermal load of the regeneration process by lowering the desorption heat and vaporization heat required for regeneration. Catalysts accelerate proton transfer processes by providing acidic sites, thereby reducing regeneration energy consumption and enhancing desorption rates. By retrofitting traditional equipment, developing new equipment, utilizing waste heat, and applying new heat pump technologies, the overall energy consumption of capture technologies can be reduced. However, practical application still faces certain challenges. Chemical desorption involves reacting calcium-based compounds with amine-rich solutions to directly sequester CO2. It can significantly reduce regeneration energy consumption, save equipment and operational costs, and the generated CaCO3 also has commercial value, offering a new pathway for sustainable development.
As one of the largest industrial sources of carbon emissions, the steel industry can reduce emissions through the implementation of carbon capture technology. Researchers have conducted simulation analyses of various carbon capture technologies for steel industry flue gas. Although these technologies have demonstrated feasibility at the laboratory and pilot scale, they are still some distance from commercialization. Therefore, commercial carbon capture in the global steel industry is still in its developmental stage, with very few operational projects currently in place. In the future, it will be necessary to further develop absorbents with high absorption rates, low regeneration energy consumption, and low costs to promote their large-scale application. At the same time, policy incentives should be used to increase investment in the research and development of low-energy carbon capture technologies, and relevant technical standards and regulations should be established to promote the implementation of CCUS technologies in the steel industry. Hence, it helps the industry achieve a balance between economic and environmental benefits during its low-carbon transition, fully unlocking its potential for sustainable development.

Funding

This work is supported by National Key R&D Program of China (No. 2022YFE0208100).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

Authors Le Ren, Tao Xie, Qianxuan Zhang, Rui Li, and Changqing Cai are employed by Capital Engineering & Research Incorporation Limited and Beijing SDL Technology Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Dziejarski, B.; Krzyżyńska, R.; Andersson, K. Current status of carbon capture, utilization, and storage technologies in the global economy: A survey of technical assessment. Fuel 2023, 342, 127776. [Google Scholar] [CrossRef]
  2. Lan, X.; Tans, P.; Thoning, K. Trends in globally-averaged CO2 Determined from NOAA Global Monitoring Laboratory Measurements. [DS/OL]. NOAA GML. 2023. Available online: https://gml.noaa.gov/ccgg/trends/global.html?doi=10.15138/9n0h-zh07 (accessed on 21 July 2025).
  3. NOAA National Centers for Environmental Information. Annual 2024 Global Climate Report [EB/OL]. 2024. Available online: https://www.ncei.noaa.gov/access/monitoring/monthly-report/global/202413 (accessed on 21 July 2025).
  4. Zhang, Z.; Wang, T.; Blunt, M.J.; Anthony, E.J.; Park, A.H.A.; Hughes, R.W.; Webley, P.A.; Yan, J. Advances in carbon capture, utilization and storage. Appl. Energy 2020, 278, 115627. [Google Scholar] [CrossRef]
  5. Yan, J.; Zhang, Z. Carbon capture, utilization and storage (CCUS). Appl. Energy 2019, 235, 1289–1299. [Google Scholar] [CrossRef]
  6. Ahmed, S.; Khan, M.K.; Kim, J. Revolutionary advancements in carbon dioxide valorization via metal-organic framework-based strategies. Carbon Capture Sci. Technol. 2025, 15, 100405. [Google Scholar] [CrossRef]
  7. Ahmed, S.; Hussain, M.S.; Khan, M.K.; Kim, J. Innovations in catalysis towards efficient electrochemical reduction of CO2 to C1 chemicals. J. Energy Chem. 2025, 107, 622–649. [Google Scholar] [CrossRef]
  8. Zhou, S.; Kosari, M.; Zeng, H.C. Boosting CO2 hydrogenation to methanol over monolayer MoS2 nanotubes by creating more strained basal planes. J. Am. Chem. Soc. 2024, 146, 10032–10043. [Google Scholar] [CrossRef] [PubMed]
  9. Zhou, S.; Ma, W.; Anjum, U.; Kosari, M.; Xi, S.; Kozlov, S.M.; Zeng, H.C. Strained few-layer MoS2 with atomic copper and selectively exposed in-plane sulfur vacancies for CO2 hydrogenation to methanol. Nat. Commun. 2023, 14, 5872. [Google Scholar] [CrossRef]
  10. Wilberforce, T.; Baroutaji, A.; Soudan, B.; Al-Alami, A.H.; Olabi, A.G. Outlook of carbon capture technology and challenges. Sci. Total Environ. 2019, 657, 56–72. [Google Scholar] [CrossRef]
  11. 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]
  12. Yadav, S.; Mondal, S.S. A review on the progress and prospects of oxy-fuel carbon capture and sequestration (CCS) technology. Fuel 2022, 308, 122057. [Google Scholar] [CrossRef]
  13. Shyam, A.; Ahmed, K.R.A.; Kumar, J.P.N.; Iniyan, S.; Goic, R. Path of carbon dioxide capture technologies: An overview. Next Sustain. 2025, 6, 100118. [Google Scholar] [CrossRef]
  14. Akinmoladun, A.; Tomomewo, O.S. Advances and future perspectives in post-combustion carbon capture technology using chemical absorption process: A review. Carbon Capture Sci. Technol. 2025, 16, 100461. [Google Scholar] [CrossRef]
  15. Hanson, E.; Nwakile, C.; Hammed, V.O. Carbon capture, utilization, and storage (CCUS) technologies: Evaluating the effectiveness of advanced CCUS solutions for reducing CO2 emissions. Results Surf. Interfaces 2025, 18, 100381. [Google Scholar] [CrossRef]
  16. Bui, M.; Adjiman, C.S.; Bardow, A.; Anthony, E.J.; Boston, A.; Brown, S.; Fennell, P.S.; Fuss, S.; Galindo, A.; Hackett, L.A.; et al. Carbon capture and storage (CCS): The way forward. Energy Environ. Sci. 2018, 11, 1062–1176. [Google Scholar] [CrossRef]
  17. Soo, X.Y.D.; Lee, J.J.C.; Wu, W.Y.; Tao, L.; Wang, C.; Zhu, Q.; Bu, J. Advancements in CO2 capture by absorption and adsorption: A comprehensive review. J. CO2 Util. 2024, 81, 102727. [Google Scholar] [CrossRef]
  18. Zhang, G.; Liu, J.; Qian, J.; Zhang, X.; Liu, Z. Review of research progress and stability studies of amine-based biphasic absorbents for CO2 capture. J. Ind. Eng. Chem. 2024, 134, 28–50. [Google Scholar] [CrossRef]
  19. Wang, N.; Zhou, J.; Ren, J. Recent advances in CO2 capture and utilization: From the perspective of process integration and optimization. Renew. Sustain. Energy Rev. 2025, 216, 115688. [Google Scholar] [CrossRef]
  20. Lei, T.; Wang, D.; Yu, X.; Ma, S.; Zhao, W.; Cui, C.; Meng, J.; Tao, S.; Guan, D. Global iron and steel plant CO2 emissions and carbon-neutrality pathways. Nature 2023, 622, 514–520. [Google Scholar] [CrossRef]
  21. Wang, Y.; Liu, J.; Tang, X.; Wang, Y.; An, H.; Yi, H. Decarbonization pathways of China’s iron and steel industry toward carbon neutrality. Resour. Conserv. Recycl. 2023, 194, 106994. [Google Scholar] [CrossRef]
  22. Andrade, C.; Desport, L.; Selosse, S. Net-negative emission opportunities for the iron and steel industry on a global scale. Appl. Energy 2024, 358, 122566. [Google Scholar] [CrossRef]
  23. Perpiñán, J.; Peña, B.; Bailera, M.; Eveloy, V.; Kannan, P.; Raj, A.; Lisbona, P.; Romeo, L.M. Integration of carbon capture technologies in blast furnace based steel making: A comprehensive and systematic review. Fuel 2023, 336, 127074. [Google Scholar] [CrossRef]
  24. Ren, L.; Zhou, S.; Peng, T.; Ou, X. A review of CO2 emissions reduction technologies and low-carbon development in the iron and steel industry focusing on China. Renew. Sustain. Energy Rev. 2021, 143, 110846. [Google Scholar] [CrossRef]
  25. Vaziri, P.; Reza Rasaei, M.; Seyfoori, S.; Zamani, S.; Mahmoodi, M.; Sedaee, B. Advancing carbon capture technologies in CCS: A comprehensive review of pre-combustion processes. Gas Sci. Eng. 2024, 131, 205481. [Google Scholar] [CrossRef]
  26. Olabi, A.G.; Wilberforce, T.; Elsaid, K.; Sayed, E.T.; Maghrabie, H.M.; Abdelkareem, M.A. Large scale application of carbon capture to process industries—A review. J. Clean. Prod. 2022, 362, 132300. [Google Scholar] [CrossRef]
  27. Xie, Y.; Xia, Z.; Li, C.; Mou, M.; Hu, X.; Chen, W.; Xu, X.; Xia, Y.; Liu, Z.; Fang, M. Pilot test and process modification of CO2 chemical absorption for top gas recycling-oxygen blast furnace. Sep. Purif. Technol. 2025, 370, 133148. [Google Scholar] [CrossRef]
  28. Ochedi, F.O.; Yu, J.; Yu, H.; Liu, Y.; Hussain, A. Carbon dioxide capture using liquid absorption methods: A review. Environ. Chem. Lett. 2021, 19, 77–109. [Google Scholar] [CrossRef]
  29. Yang, S.; Zhang, L.; Song, D. Conceptual design, optimization and thermodynamic analysis of a CO2 capture process based on rectisol. Energy 2022, 244, 122566. [Google Scholar] [CrossRef]
  30. Emmanuel, O.; Rozina Ezeji, T.C. Advances in carbon dioxide capture and conversion technologies: Industrial integration for sustainable chemical production. Next Sustain. 2025, 6, 100108. [Google Scholar] [CrossRef]
  31. Wang, J.; Shen, Y.; Zhang, D.; Tang, Z.; Li, W. Integrated vacuum pressure swing adsorption and rectisol process for CO2 capture from underground coal gasification syngas. Chin. J. Chem. Eng. 2023, 57, 265–279. [Google Scholar] [CrossRef]
  32. Wang, N.; Lu, S.; Liu, L.; Lang, J.; Liu, M.; Sun, M.; Kang, G. Research progress of catalytic regeneration for energy-efficient CO2 capture in amine absorption system. Chem. Ind. Eng. Prog. 2025, 44, 445–464. [Google Scholar]
  33. Wang, Q.; Gong, D.; Zhang, Z.; Zhu, J.; Luo, Q. Experimental study on the cyclic mineralization of CO2 enriched phase after absorption by a novel biphasic absorbent composed of DEEA and AEP. Sci. Rep. 2024, 14, 26759. [Google Scholar] [CrossRef] [PubMed]
  34. Kim Smang Léonard, G. CO2 capture technologies and shortcut cost correlations for different inlet CO2 concentrations and flow rates. part 1: Chemical absorption. Int. J. Greenh. Gas Control 2025, 145, 104391. [Google Scholar]
  35. Du, J.; Yang, W.; Xu, L.; Bei, L.; Lei, S.; Li, W.; Liu, H.; Wang, B.; Sun, L. Review on post-combustion CO2 capture by amine blended solvents and aqueous ammonia. Chem. Eng. J. 2024, 488, 150954. [Google Scholar] [CrossRef]
  36. Kim, J.; Kim, K.; Lim, H.; Kang, J.H.; Park, H.S.; Park, J.; Song, H. Structural investigation of aqueous amine solutions for CO2 capture: CO2 loading, cyclic capacity, absorption–desorption rate, and pKa. J. Environ. Chem. Eng. 2024, 12, 112664. [Google Scholar] [CrossRef]
  37. Ahmadi, M.; Bahmanzadegan, F.; Qasemnazhand, M.; Ghaemi, A.; Ramezanipour Penchah, H. Experimental, RSM modelling, and DFT simulation of CO2 adsorption on modified activated carbon with LiOH. Sci. Rep. 2024, 14, 13595. [Google Scholar] [CrossRef]
  38. Siegelman, R.L.; Kim, E.J.; Long, J.R. Porous materials for carbon dioxide separations. Nat. Mater. 2021, 20, 1060–1072. [Google Scholar] [CrossRef]
  39. Chang, X.L.; Yan, T.; Pan, W.G.; Wang, L.W. Designing and screening metal-organic frameworks for enhancing CO2 capture and separation performance. Coord. Chem. Rev. 2025, 543, 216929. [Google Scholar] [CrossRef]
  40. Li, J.; Xu, Z.; Yang, X. A theoretical simulation of carbon dioxide adsorption capture on amine-functionalized graphene oxides. J. Environ. Chem. Eng. 2024, 12, 114165. [Google Scholar] [CrossRef]
  41. Liang, W.; Zhang, Y.; Yang, L.; Liu, J. Adsorption performance of amine-functionalized red mud-based adsorbent for CO2 capture. J. Environ. Manag. 2025, 383, 125334. [Google Scholar] [CrossRef]
  42. Zhu, S.; Zhao, B.; Su, Y. Multiple-amine functionalized γ–Al2O3 for post-combustion CO2 capture: Performance, mechanism, and kinetics. Fuel 2025, 380, 133186. [Google Scholar] [CrossRef]
  43. Zhao, S.; Zhang, Y.; Li, L.; Feng, J.; Qiu, W.; Ning, Y.; Huang, Z.; Lin, H. Degradation of amine-functionalized adsorbents in carbon capture and direct air capture applications: Mechanism and solutions. Sep. Purif. Technol. 2025, 354, 129586. [Google Scholar] [CrossRef]
  44. Yan, H.; Zhang, G.; Liu, J.; Li, G.; Zhao, Y.; Wang, Y.; Wu, C.; Zhang, Y.; Xu, Y. Investigation of CO2 adsorption performance of amine impregnated adsorbents using amine-support matching strategies. Sep. Purif. Technol. 2023, 310, 123178. [Google Scholar] [CrossRef]
  45. Nguyen, T.S.; Dogan, N.A.; Lim, H.; Yavuz, C.T. Amine chemistry of porous CO2 adsorbents. Acc. Chem. Res. 2023, 56, 2642–2652. [Google Scholar] [CrossRef] [PubMed]
  46. Siagian, U.W.R.; Raksajati, A.; Himma, N.F.; Khoiruddin, K.; Wenten, I.G. Membrane-based carbon capture technologies: Membrane gas separation vs. membrane contactor. J. Nat. Gas Sci. Eng. 2019, 67, 172–195. [Google Scholar] [CrossRef]
  47. Niu, Z.; He, N.; Yao, Y.; Ma, A.; Zhang, E.; Cheng, L.; Li, Y.; Lu, X. Mixed matrix membranes for gas separations: A review. Chem. Eng. J. 2024, 494, 152912. [Google Scholar] [CrossRef]
  48. Xie, K.; Fu, Q.; Qiao, G.G.; Webley, P.A. Recent progress on fabrication methods of polymeric thin film gas separation membranes for CO2 capture. J. Membr. Sci. 2019, 572, 38–60. [Google Scholar] [CrossRef]
  49. Wang, Y.; Wang, J.; Sun, L.; Ma, X. Progress in polymers of intrinsic microporosity for post-combustion carbon capture. Sep. Purif. Technol. 2025, 363, 132003. [Google Scholar] [CrossRef]
  50. Han, Y.; Ho, W.S.W. Polymeric membranes for CO2 separation and capture. J. Membr. Sci. 2021, 628, 119244. [Google Scholar] [CrossRef]
  51. Prasetya, N.; Himma, N.F.; Sutrisna, P.D.; Wenten, I.G.; Ladewig, B.P. A review on emerging organic-containing microporous material membranes for carbon capture and separation. Chem. Eng. J. 2020, 391, 123575. [Google Scholar] [CrossRef]
  52. Gkotsis, P.; Peleka, E.; Zouboulis, A. Membrane-based technologies for post-combustion CO2 capture from flue gases: Recent progress in commonly employed membrane materials. Membranes 2023, 13, 898. [Google Scholar] [CrossRef]
  53. Wang, J.; Li, L.; Li, X.; Zhang, J. Constructing double-faced CO2-capture domains by sandwich-like fillers in membranes for efficient CO2 separation. Chem. Eng. Sci. 2024, 283, 119374. [Google Scholar] [CrossRef]
  54. Cao, H.; Cao, X.; Wang, Z.; Ding, G.; Li, H.; Bian, J. Advancing cryogenic carbon capture technology through understanding the CO2 frosting crystallization mechanism. Fuel 2025, 384, 134006. [Google Scholar] [CrossRef]
  55. Asgharian, H.; Iov, F.; Araya, S.S.; Pedersen, T.H.; Nielsen, M.P.; Baniasadi, E.; Liso, V. A review on process modeling and simulation of cryogenic carbon capture for post-combustion treatment. Energies 2023, 16, 1855. [Google Scholar] [CrossRef]
  56. Du, S.; Qiu, G.; Li, Q.; Jiang, L.; Cai, W. Cryogenic desublimation carbon capture technology: Mechanisms and applications. Sep. Purif. Technol. 2025, 375, 133677. [Google Scholar] [CrossRef]
  57. Shen, M.; Tong, L.; Yin, S.; Liu, C.; Wang, L.; Feng, W.; Ding, Y. Cryogenic technology progress for CO2 capture under carbon neutrality goals: A review. Sep. Purif. Technol. 2022, 299, 121734. [Google Scholar] [CrossRef]
  58. De, D.K.; Oduniyi, I.A.; Sam, A.A.; Aneesh, A.M.; Akinmeji, S. Modelling carbon capture from power plants with low energy and water consumption using a novel cryogenic technology. Appl. Therm. Eng. 2024, 257, 124315. [Google Scholar] [CrossRef]
  59. Song, C.; Run, L.; Liu, Q.; Deng, S.; Li, H.; Sun, J. Membrane-cryogenic hybrid CO2 capture—A review. Carbon Capture Sci. Technol. 2024, 13, 100249. [Google Scholar] [CrossRef]
  60. Mondal, M.K.; Balsora, H.K.; Varshney, P. Progress and trends in CO2 capture/separation technologies: A review. Energy 2012, 46, 431–441. [Google Scholar] [CrossRef]
  61. Marques, L.M.; Mota, S.M.; Teixeira, P.; Mateus, M.; Pinheiro, C.I.C. Decarbonisation of cement industry: Calcium looping with white mud and limestone as CaO-based sorbents for industrial flue gas treatment. Chem. Eng. Process.—Process Intensif. 2025, 213, 110290. [Google Scholar] [CrossRef]
  62. Rincon Duarte, J.P.; Kriechbaumer, D.; Lachmann, B.; Tescari, S.; Fend, T.; Roeb, M.; Sattler, C. Solar calcium looping cycle for CO2 capturing in a cement plant. Definition of process parameters and reactors selection. Sol. Energy 2022, 238, 189–202. [Google Scholar] [CrossRef]
  63. Zhang, Y.; Wang, Y.; Han, K.; Zhao, J.; Wu, J.J.; Li, Y. Calcium looping for CO2 capture and thermochemical heat storage, a potential technology for carbon neutrality: A review. Green Energy Resour. 2024, 2, 100078. [Google Scholar] [CrossRef]
  64. Behtash, H.R.; Tahmasebpoor, M. Experimental and modeling insights into the kinetics of calcium looping CO2 capture process: Assessment of limestone treatment. J. Environ. Chem. Eng. 2024, 12, 114645. [Google Scholar] [CrossRef]
  65. Rong, N.; Wang, S.; Chu, C.; Guo, Z.; Liu, K.; Han, L.; Ge, L.; Shi, X.; Wang, G.; Wang, Y. Deactivated ca-based sorbent derived from calcium looping CO2 capture as a partial substitute for cement to obtain low-carbon cementitious building materials. Constr. Build. Mater. 2024, 454, 139175. [Google Scholar] [CrossRef]
  66. Alalwan, H. A, Alminshid, A.H. CO2 capturing methods: Chemical looping combustion (CLC) as a promising technique. Sci. Total Environ. 2021, 788, 147850. [Google Scholar] [CrossRef]
  67. Yan, J.; Wang, S.; Hu, B.; Zhang, H.; Qiu, L.; Liu, W.; Guo, Y.; Shen, J.; Chen, B.; Ge, X.; et al. Review of chemical looping technology for energy conservation and utilization: CO2 capture and energy cascade utilization. J. Environ. Chem. Eng. 2024, 12, 112602. [Google Scholar] [CrossRef]
  68. Saeedan, M.; Houshfar, E. Solar-driven chemical looping combustion: A pathway to low-impact carbon emission and sustainable hydrogen generation for a decarbonized energy sector. Fuel Process. Technol. 2025, 273, 108230. [Google Scholar] [CrossRef]
  69. Baskaran, D.; Saravanan, P.; Nagarajan, L.; Byun, H.S. An overview of technologies for capturing, storing, and utilizing carbon dioxide: Technology readiness, large-scale demonstration, and cost. Chem. Eng. J. 2024, 491, 151998. [Google Scholar] [CrossRef]
  70. Zuo, H.; Mao, J.; Duan, Z.; Qian, A.; Shi, Y.; Gu, Z.; Li, D.; Wei, Y.; Wang, H.; Li, K. Separation and purification of CO2 from blast furnace gas combustion by a chemical looping concept. Sep. Purif. Technol. 2025, 370, 133084. [Google Scholar] [CrossRef]
  71. Jin, B.; Xu, M.; Wang, R.; Fan, Y.; Xiang, X.; Zhao, H.; Liang, Z. Steel slag derived oxygen carriers and solid sorbents for chemical looping CO2 capture: A mini review. Results Eng. 2025, 25, 104438. [Google Scholar] [CrossRef]
  72. Vaz, S.; De Souza, A.P.R.; Baeta, B.E.L. Technologies for carbon dioxide capture: A review applied to energy sectors. Clean. Eng. Technol. 2022, 8, 100456. [Google Scholar] [CrossRef]
  73. Komkhum, T.; Sema, T.; Rehman, Z.U.; In-na, P. Carbon dioxide removal from triethanolamine solution using living microalgae-loofah biocomposites. Sci. Rep. 2025, 15, 7247. [Google Scholar] [CrossRef]
  74. Meng, F.; Meng, Y.; Ju, T.; Han, S.; Lin, L.; Jiang, J. Research progress of aqueous amine solution for CO2 capture: A review. Renew. Sustain. Energy Rev. 2022, 168, 112902. [Google Scholar] [CrossRef]
  75. Pérez-Calvo, J.F.; Sutter, D.; Gazzani, M.; Mazzotti, M. Advanced configurations for post-combustion CO2 capture processes using an aqueous ammonia solution as absorbent. Sep. Purif. Technol. 2021, 274, 118959. [Google Scholar] [CrossRef]
  76. Zhu, Y.; Zhang, B.; Deng, C.; Lei, Q.; Chen, H. Efficient desorption and energy-saving carbon capture using a nonaqueous primary alkanolamine-based biphasic absorbent. Sep. Purif. Technol. 2025, 356, 129855. [Google Scholar] [CrossRef]
  77. Singto, S.; Supap, T.; Idem, R.; Tontiwachwuthikul, P.; Tantayanon, S.; Al-Marri, M.J.; Benamor, A. Synthesis of new amines for enhanced carbon dioxide (CO2) capture performance: The effect of chemical structure on equilibrium solubility, cyclic capacity, kinetics of absorption and regeneration, and heats of absorption and regeneration. Sep. Purif. Technol. 2016, 167, 97–107. [Google Scholar] [CrossRef]
  78. Narimani, M.; Amjad-Iranagh, S.; Modarress, H. Performance of tertiary amines as the absorbents for CO2 capture: Quantum mechanics and molecular dynamics studies. J. Nat. Gas Sci. Eng. 2017, 47, 154–166. [Google Scholar] [CrossRef]
  79. Xiao, L.; Qiu, Z.; Feng, S.; Duan, X.; Zhao, Z.; Liu, Y.; Ma, L. Carbon dioxide absorption and desorption experiments based on MDEA. Chem. Eng. Process.—Process Intensif. 2024, 204, 109931. [Google Scholar] [CrossRef]
  80. 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]
  81. Muchan, P.; Saiwan, C.; Narku-Tetteh, J.; Idem, R.; Supap, T.; Tontiwachwuthikul, P. Screening tests of aqueous alkanolamine solutions based on primary, secondary, and tertiary structure for blended aqueous amine solution selection in post combustion CO2 capture. Chem. Eng. Sci. 2017, 170, 574–582. [Google Scholar] [CrossRef]
  82. Fonck, F.; Karlsson, H.K.; Antonopoulou, I.; Svensson, H. Evaluation of enhanced absorption of carbon dioxide using carbonic anhydrase in aqueous solutions of 2-amino-2-methyl-1-propanol. Clean. Eng. Technol. 2025, 25, 100918. [Google Scholar] [CrossRef]
  83. Sung, H.; Han, S.C.; Jeon, P.R.; Kim, K.M. The effects of blending diisopropanolamine and 2-amino-2-methyl-1-propanol for carbon dioxide capture process: Comparative study using thermodynamic modeling and artificial neural network predictions. J. Clean. Prod. 2025, 506, 145474. [Google Scholar] [CrossRef]
  84. Nwaoha, C.; Idem, R.; Supap, T.; Saiwan, C.; Tontiwachwuthikul, P.; Rongwong, W.; Al-Marri, M.J.; Benamor, A. Heat duty, heat of absorption, sensible heat and heat of vaporization of 2–amino–2–methyl–1–propanol (AMP), piperazine (PZ) and monoethanolamine (MEA) tri–solvent blend for carbon dioxide (CO2) capture. Chem. Eng. Sci. 2017, 170, 26–35. [Google Scholar] [CrossRef]
  85. Xiang, J.; Wei, D.; Mao, W.; Liu, T.; Luo, Q.; Huang, Y.; Liang, Z.; Luo, X. Comprehensive kinetic study of carbon dioxide absorption in blended tertiary/secondary amine solutions: Experiments and simulations. Sep. Purif. Technol. 2024, 330, 125310. [Google Scholar] [CrossRef]
  86. Zhang, W.; Xu, Y.; Wang, Q. Progress of research on regeneration of rich alkanolamine solution with low energy consumption. Chem. Ind. Eng. Prog. 2021, 40, 4497–4507. [Google Scholar]
  87. Lu, G.; Farrukh, S.; Fan, X. Research progress of non-aqueous absorbents for carbon dioxide capture with low energy consumption: A review. Fuel 2025, 391, 134740. [Google Scholar] [CrossRef]
  88. Bihong, L.; Kexuan, Y.; Xiaobin, Z.; Zuoming, Z.; Guohua, J. 2-amino-2-methyl-1-propanol based non-aqueous absorbent for energy-efficient and non-corrosive carbon dioxide capture. Appl. Energy 2020, 264, 114703. [Google Scholar] [CrossRef]
  89. Chen, M.; Li, M.; Zhang, F.; Hu, X.; Wu, Y. Fast and efficient CO2 absorption in non-aqueous tertiary amines promoted by ethylene glycol. Energy Fuels 2022, 36, 4830–4836. [Google Scholar] [CrossRef]
  90. Guo, H.; Li, C.; Shi, X.; Li, H.; Shen, S. Nonaqueous amine-based absorbents for energy efficient CO2 capture. Appl. Energy 2019, 239, 725–734. [Google Scholar] [CrossRef]
  91. Wang, L.; Wang, X.; Ma, J.; Meng, F.; Li, K.; Li, M.; Wang, R.; Wang, H.; Gao, H.; Zhao, K.; et al. Performance and mechanisms of CO2 capture by water-lean 3-diethylaminopropylamine (DEAPA) based absorbents. Energy 2025, 320, 135232. [Google Scholar] [CrossRef]
  92. Krishnan, A.; Gopinath, K.P.; Vo, D.V.N.; Malolan, R.; Nagarajan, V.M.; Arun, J. Ionic liquids, deep eutectic solvents and liquid polymers as green solvents in carbon capture technologies: A review. Environ. Chem. Lett. 2020, 18, 2031–2054. [Google Scholar] [CrossRef]
  93. 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. [Google Scholar] [CrossRef]
  94. Cao, L.; Gao, J.; Zeng, S.; Dong, H.; Gao, H.; Zhang, X.; Huang, J. Feasible ionic liquid-amine hybrid solvents for carbon dioxide capture. Int. J. Greenh. Gas Control 2017, 66, 120–128. [Google Scholar] [CrossRef]
  95. Wu, Y.; Xu, J.; Mumford, K.; Stevens, G.W.; Fei, W.; Wang, Y. Recent advances in carbon dioxide capture and utilization with amines and ionic liquids. Green Chem. Eng. 2020, 1, 16–32. [Google Scholar] [CrossRef]
  96. Zafar, A.; Matuszek, K.; MacFarlane, D.R.; Zhang, X. Recent progress and future perspectives of ionic liquid-based carbon dioxide capture and conversion. Green Energy Environ. 2025, 10, 1097–1129. [Google Scholar] [CrossRef]
  97. Zhang, Y.; Ji, X.; Xie, Y.; Lu, X. Screening of conventional ionic liquids for carbon dioxide capture and separation. Appl. Energy 2016, 162, 1160–1170. [Google Scholar] [CrossRef]
  98. Lei, Y.; Du, L.; Hu, S.; Kuang, Y.; Liu, X.; Fang, H.; Chen, Y. Ionic liquid-ethanol mixed solvent design for the post-combustion carbon capture. Sep. Purif. Technol. 2025, 354, 129318. [Google Scholar] [CrossRef]
  99. Ding, X.; Li, X.; Pimentel, J.; Zhou, T. CO2 capture using blended amine—Ionic liquid solvents: Thermodynamic modeling and process optimization. Sep. Purif. Technol. 2025, 362, 131649. [Google Scholar] [CrossRef]
  100. Min, G.H.; Park, H.J.; Son, Y.S.; Shin, D.; Baek, I.H.; Lee, S.; Nam, S.C. Effective amine solvent regeneration over NiO-modified SBA-15 catalysts for energy-saving CO2 capture. J. CO2 Util. 2024, 85, 102873. [Google Scholar] [CrossRef]
  101. Guo, Y.; Su, Z.; Mao, W.; Xie, H.; Tan, Z.; Gong, Z.; Zhang, X. Catalytic behavior of aluminum-modified SAPO-34 catalysts for reducing the energy penalty of CO2-rich amine solutions regeneration. Sep. Purif. Technol. 2025, 353, 128327. [Google Scholar] [CrossRef]
  102. Zhang, R.; Li, T.; Zhang, Y.; Ha, J.; Xiao, Y.; Li, C.; Zhang, X.; Luo, H. CuO modified KIT-6 as a high-efficiency catalyst for energy-efficient amine solvent regeneration. Sep. Purif. Technol. 2022, 300, 121702. [Google Scholar] [CrossRef]
  103. Wei, K.; Xing, L.; Li, Y.; Xu, T.; Li, Q.; Wang, L. Heteropolyacid modified cerium-based MOFs catalyst for amine solution regeneration in CO2 capture. Sep. Purif. Technol. 2022, 293, 121144. [Google Scholar] [CrossRef]
  104. Ai, S.; Xiang, Y.; Chen, X.; Zhang, Y.; Zhang, Y.; Lv, K.; Chen, X.; Yu, D.; Ge, B.; Huang, F. ZIF-8 and ZIF-67 as catalysts for promoting carbon dioxide capture based on monoethanolamine solution. Sep. Purif. Technol. 2025, 358, 130389. [Google Scholar] [CrossRef]
  105. Tan, Z.; Zhang, S.; Zhao, F.; Zhang, R.; Tang, F.; You, K.; Luo, H.; Zhang, X. SnO2/ATP catalyst enabling energy-efficient and green amine-based CO2 capture. Chem. Eng. J. 2023, 453, 139801. [Google Scholar] [CrossRef]
  106. Zhang, X.; Huang, Y.; Yang, J.; Gao, H.; Huang, Y.; Luo, X.; Liang, Z.; Tontiwachwuthikul, P. Amine-based CO2 capture aided by acid-basic bifunctional catalyst: Advancement of amine regeneration using metal modified MCM-41. Chem. Eng. J. 2020, 383, 123077. [Google Scholar] [CrossRef]
  107. Gao, H.; Huang, Y.; Zhang, X.; Bairq, Z.A.S.; Huang, Y.; Tontiwachwuthikul, P.; Liang, Z. Catalytic performance and mechanism of SO42−/ZrO2/SBA-15 catalyst for CO2 desorption in CO2-loaded monoethanolamine solution. Appl. Energy 2020, 259, 114179. [Google Scholar] [CrossRef]
  108. Yang, X.; Zhou, R.; Liu, G.; Hussain, S.; Li, Q. Reduce energy consumption for organic amine regeneration by MnOOH/HZSM-5 catalysts. Chem. Eng. J. 2024, 493, 152564. [Google Scholar] [CrossRef]
  109. Waseem, M.; Al-Marzouqi, M.; Ghasem, N. Enhancing regeneration energy efficiency of CO2-rich amine solution with a novel tri-composite catalyst. J. CO2 Util. 2024, 82, 102764. [Google Scholar] [CrossRef]
  110. Oh, S.Y.; Binns, M.; Cho, H.; Kim, J.K. Energy minimization of MEA-based CO2 capture process. Appl. Energy 2016, 169, 353–362. [Google Scholar] [CrossRef]
  111. Jung, J.; Jeong, Y.S.; Lee, U.; Lim, Y.; Han, C. New configuration of the CO2 capture process using aqueous monoethanolamine for coal-fired power plants. Ind. Eng. Chem. Res. 2015, 54, 3865–3878. [Google Scholar] [CrossRef]
  112. Zhang, Z.; Hong, S.H.; Lee, C.H. Role and impact of wash columns on the performance of chemical absorption-based CO2 capture process for blast furnace gas in iron and steel industries. Energy 2023, 271, 127020. [Google Scholar] [CrossRef]
  113. Lee, W.H.; Kim, G.; Kim, J.O. Performance of a novel CO2 absorption and reverse electrodialysis system with different amine solutions for simultaneous energy and bicarbonate recovery. Energy Convers. Manag. 2024, 314, 118715. [Google Scholar] [CrossRef]
  114. Tu, T.; Wu, T.; Zhang, X.; Liang, F.; Ji, L.; Yan, S. Coupling gas-assisted stripping and waste heat recovery for energy-efficient carbon capture process. Green Carbon, 2025, in press. [CrossRef]
  115. Lei, T.; Liang, Y.; Zhu, Y.; Ye, K.; Wang, J.; Liang, Y. A novel post-combustion CO2 capture process for natural gas combined cycle power plant based on waste energy utilization and absorption heat transformer. Energy 2025, 316, 134491. [Google Scholar] [CrossRef]
  116. Ji, L.; Yu, H.; Li, K.; Yu, B.; Grigore, M.; Yang, Q.; Wang, X.; Chen, Z.; Zeng, M.; Zhao, S. Integrated absorption-mineralisation for low-energy CO2 capture and sequestration. Appl. Energy 2018, 225, 356–366. [Google Scholar] [CrossRef]
  117. Arti, M.; Youn, M.H.; Park, K.T.; Kim, H.J.; Kim, Y.E.; Jeong, S.K. Single process for CO2 capture and mineralization in various alkanolamines using calcium chloride. Energy Fuels 2017, 31, 763–769. [Google Scholar] [CrossRef]
  118. Zhang, W.; Deng, Z.; Wang, Q.; Li, J.; Qiu, X. Experimental research on chemical desorption based on CO2-rich absorption solutions. Int. J. Greenh. Gas Control 2021, 109, 103356. [Google Scholar] [CrossRef]
  119. Shahid, K.; Nguyen, H.; Unluer, C.; Kinnunen, P. Development of alternative for gypsum-based plaster using magnesium carbonates from carbon capture and utilization process. Constr. Build. Mater. 2024, 447, 137999. [Google Scholar] [CrossRef]
  120. Park, S. CO2 reduction-conversion to precipitates and morphological control through the application of the mineral carbonation mechanism. Energy 2018, 153, 413–421. [Google Scholar] [CrossRef]
  121. Chen, T.L.; Jiang, W.; Shen, A.L.; Chen, Y.H.; Pan, S.Y.; Chiang, P.C. CO2 mineralization and utilization using various calcium-containing wastewater and refining slag via a high-gravity carbonation process. Ind. Eng. Chem. Res. 2020, 59, 7140–7150. [Google Scholar] [CrossRef]
  122. Kang, J.M.; Murnandari, A.; Youn, M.H.; Lee, W.; Park, K.T.; Kim, Y.E.; Kim, H.J.; Kang, S.P.; Lee, J.H.; Jeong, S.K. Energy-efficient chemical regeneration of AMP using calcium hydroxide for operating carbon dioxide capture process. Chem. Eng. J. 2018, 335, 338–344. [Google Scholar] [CrossRef]
  123. Chen, K.; Han, S.; Meng, F.; Lin, L.; Li, J.; Gao, Y.; Qin, W.; Hu, E.; Jiang, J. Diisobutylamine mediated CO2 mineralization and CaCO3 production from municipal solid waste incineration fly ash as raw ingredient and regeneration reagent. Chem. Eng. J. 2024, 481, 148392. [Google Scholar] [CrossRef]
  124. Lin, X.; Zhang, Y.; Liu, H.; Boczkaj, G.; Cao, Y.; Wang, C. Carbon dioxide sequestration by industrial wastes through mineral carbonation: Current status and perspectives. J. Clean. Prod. 2024, 434, 140258. [Google Scholar] [CrossRef]
  125. Wang, Y.; Song, L.; Ma, K.; Liu, C.; Tang, S.; Yan, Z.; Yue, H.; Liang, B. An integrated absorption–mineralization process for CO2 capture and sequestration: Reaction mechanism, recycling stability, and energy evaluation. ACS Sustain. Chem. Eng. 2021, 9, 16577–16587. [Google Scholar] [CrossRef]
  126. Cheng, S.; Zheng, Y.; Li, G.; Gao, J.; Li, R.; Yue, T. Research progress on phase change absorbents for CO2 capture in industrial flue gas: Principles and application prospects. Sep. Purif. Technol. 2025, 354, 129296. [Google Scholar] [CrossRef]
  127. An, S.; Xu, T.; Xing, L.; Yu, G.; Zhang, R.; Liu, J.; Aierken, A.; Dai, Q.; Wang, L. Recent progress and prospects in solid acid-catalyzed CO2 desorption from amine-rich liquid. Gas Sci. Eng. 2023, 120, 205152. [Google Scholar] [CrossRef]
  128. Khallaghi, N.; Abbas, S.Z.; Manzolini, G.; De Coninck, E.; Spallina, V. Techno-economic assessment of blast furnace gas pre-combustion decarbonisation integrated with the power generation. Energy Convers. Manag. 2022, 255, 115252. [Google Scholar] [CrossRef]
  129. Yang, Y.; Du, T.; Li, Y.; Yue, Q.; Wang, H.; Liu, L.; Che, S.; Wang, Y. Techno-economic assessment and exergy analysis of iron and steel plant coupled MEA- CO2 capture process. J. Clean. Prod. 2023, 416, 137976. [Google Scholar] [CrossRef]
  130. Fang, M.; Xie, Y.; Shao, Y.; Xu, X.; Xia, Y.; Mou, M.; Hu, X.; Wang, T.; Liu, Z. Development of carbon capture absorbents for top gas recycling-oxygen blast furnace in the steel industry. Sep. Purif. Technol. 2025, 355, 129616. [Google Scholar] [CrossRef]
  131. Yun, S.; Jang, M.G.; Kim, J.K. Techno-economic assessment and comparison of absorption and membrane CO2 capture processes for iron and steel industry. Energy 2021, 229, 120778. [Google Scholar] [CrossRef]
  132. Cormos, C.C. Evaluation of reactive absorption and adsorption systems for post-combustion CO2 capture applied to iron and steel industry. Appl. Therm. Eng. 2016, 105, 56–64. [Google Scholar] [CrossRef]
  133. Saeed, I.M.; Alaba, P.; Mazari, S.A.; Basirun, W.J.; Lee, V.S.; Sabzoi, N. Opportunities and challenges in the development of monoethanolamine and its blends for post-combustion CO2 capture. Int. J. Greenh. Gas Control 2018, 79, 212–233. [Google Scholar] [CrossRef]
  134. Feng, C.; Zhu, R.; Wei, G.; Dong, K.; Dong, J. Typical case of carbon capture and utilization in Chinese iron and steel enterprises: CO2 emission analysis. J. Clean. Prod. 2022, 363, 132528. [Google Scholar] [CrossRef]
  135. Pan, S.Y.; Chen, Y.H.; Fan, L.S.; Kim, H.; Gao, X.; Ling, T.C.; Chiang, P.C.; Pei, S.L.; Gu, G. CO2 mineralization and utilization by alkaline solid wastes for potential carbon reduction. Nat. Sustain. 2020, 3, 399–405. [Google Scholar] [CrossRef]
  136. Zhang, Y.; Ying, Y.; Xing, L.; Zhan, G.; Deng, Y.; Chen, Z.; Li, J. Carbon dioxide reduction through mineral carbonation by steel slag. J. Environ. Sci. 2025, 152, 664–684. [Google Scholar] [CrossRef]
  137. Global CCS Institute. The Global Status of CCS: 2024; Global CCS Institute: Melbourne, Australia, 2024. [Google Scholar]
  138. Dong, Z.; Que, L.; Li, W.; Ren, Q.; Wang, C.; Li, S.; Lv, B.; Jing, G.; Shen, H. Evaluation of 3-dibutylamino-propylamine aqueous solution for CO2 capture: Promoting the energy-saving regeneration through self-extraction regulation. Chem. Eng. J. 2024, 484, 149567. [Google Scholar] [CrossRef]
  139. Wang, Z.; Han, Z.; Liu, D.; Yang, X.; Zhou, Z.; Wu, X.; Lu, S. Investigation on CO2 capture performance of low-viscosity diamine non-aqueous absorbents with N-methyldiethanolamine regulation. Sep. Purif. Technol. 2025, 357, 130150. [Google Scholar] [CrossRef]
  140. Zhang, Y.; Dong, J.; Ning, P.; Wang, L.; Wang, J.; Ma, Y.; Wang, X. Investigation of CO2 capture performance of polyamine/organic alcohol ether non-aqueous absorbent regulated by ethylene glycol. J. Environ. Chem. Eng. 2024, 12, 113694. [Google Scholar] [CrossRef]
  141. Shen, Y.; Jiang, C.; Zhang, S.; Chen, J.; Wang, L.; Chen, J. Biphasic solvent for CO2 capture: Amine property-performance and heat duty relationship. Appl. Energy 2018, 230, 726–733. [Google Scholar] [CrossRef]
  142. Hu, Y.; Wang, Q.; Hu, D.; Zhang, Y.; Furqan, M.; Lu, S. Experimental study on CO2 capture by MEA/n-butanol/H2O phase change absorbent. RSC Adv. 2024, 14, 3146–3157. [Google Scholar] [CrossRef]
  143. Jia, R.Q.; Xu, Y.H.; Zhang, J.J.; Zhang, L.L.; Chu, G.W.; Chen, J.F. A novel phase change absorbent with ionic liquid as promoter for low energy-consuming CO2 capture. Sep. Purif. Technol. 2023, 315, 123740. [Google Scholar] [CrossRef]
  144. Zhong, X.; Kong, W.; Dong, Z.; Yang, K.; Song, T.; Wang, T.; Fang, M.; Li, W.; Li, S. A novel ILs biphasic absorbent with low regeneration energy consumption for CO2 capture: Screening of phase separation regulators and mechanism study. Chem. Eng. J. 2024, 493, 152454. [Google Scholar] [CrossRef]
  145. Aromada, S.A.; Eldrup, N.H.; Øi, L.E. Technoeconomic evaluation of combined rich and lean vapour compression configuration for CO2 capture from a cement plant. Int. J. Greenh. Gas Control 2023, 127, 103932. [Google Scholar] [CrossRef]
  146. Ye, Y.; Zhao, X.; Chen, J.; Fang, M. Pilot-scale experimental study of a new high-loading absorbent for capturing CO2 from flue gas. Processes 2022, 10, 599. [Google Scholar] [CrossRef]
  147. Artanto, Y.; Jansen, J.; Pearson, P.; Puxty, G.; Cottrell, A.; Meuleman, E.; Feron, P. Pilot-scale evaluation of AMP/PZ to capture CO2 from flue gas of an australian brown coal–fired power station. Int. J. Greenh. Gas Control 2014, 20, 189–195. [Google Scholar] [CrossRef]
  148. Rabensteiner, M.; Kinger, G.; Koller, M.; Gronald, G.; Unterberger, S.; Hochenauer, C. Investigation of the suitability of aqueous sodium glycinate as a solvent for post combustion carbon dioxide capture on the basis of pilot plant studies and screening methods. Int. J. Greenh. Gas Control 2014, 29, 1–15. [Google Scholar] [CrossRef]
  149. Vega, F.; Baena-Moreno, F.M.; Gallego Fernández, L.M.; Portillo, E.; Navarrete, B.; Zhang, Z. Current status of CO2 chemical absorption research applied to CCS: Towards full deployment at industrial scale. Appl. Energy 2020, 260, 114313. [Google Scholar] [CrossRef]
  150. Zhang, T.; Zhang, M.; Jin, L.; Xu, M.; Li, J. Advancing carbon capture in hard-to-abate industries: Technology, cost, and policy insights. Clean Technol. Environ. Policy 2024, 26, 2077–2094. [Google Scholar] [CrossRef]
  151. Colombe, C.; Leibowicz, B.D.; Mendoza, B.R. The effects of policy uncertainty and risk aversion on carbon capture, utilization, and storage investments. Energy Policy 2024, 192, 114212. [Google Scholar] [CrossRef]
  152. Song, D.; Jiang, T.; Rao, C. Review of policy framework for the development of carbon capture, utilization and storage in China. Int. J. Environ. Res. Public Health 2022, 19, 16853. [Google Scholar] [CrossRef]
Figure 1. Research framework of this paper.
Figure 1. Research framework of this paper.
Sustainability 17 09796 g001
Figure 2. Ammine absorption CO2 process diagram [34].
Figure 2. Ammine absorption CO2 process diagram [34].
Sustainability 17 09796 g002
Figure 3. CO2 adsorption reaction of amine-functionalized adsorbents [43].
Figure 3. CO2 adsorption reaction of amine-functionalized adsorbents [43].
Sustainability 17 09796 g003
Figure 4. The membrane separation technology reaction process diagram [46,48].
Figure 4. The membrane separation technology reaction process diagram [46,48].
Sustainability 17 09796 g004
Figure 5. Cryogenic process diagram [58,59].
Figure 5. Cryogenic process diagram [58,59].
Sustainability 17 09796 g005
Figure 6. Calcium looping process diagram [65].
Figure 6. Calcium looping process diagram [65].
Sustainability 17 09796 g006
Figure 7. CLC reaction process [19].
Figure 7. CLC reaction process [19].
Sustainability 17 09796 g007
Figure 8. Steel slag leaching-mineralization-regeneration process.
Figure 8. Steel slag leaching-mineralization-regeneration process.
Sustainability 17 09796 g008
Table 1. Comparison of various carbon capture technologies [1,19,72].
Table 1. Comparison of various carbon capture technologies [1,19,72].
MethodAdvantagesDisadvantages/Challenges
AbsorptionMost mature technology
Fast absorption rate
Simple, suitable for low CO2 partial pressure environments
High regeneration energy consumption
Equipment corrosion
Absorbent volatilization
Environmental impact caused by solvent degradation
AdsorptionEasy to use and maintain
Adsorbents are reusable
Low regeneration energy consumption
No corrosion issues
Low CO2 selectivity and low separation efficiency
High cost of high-efficiency adsorbents
Adsorbent wear, resulting in poor durability
Membrane separation technologyLow energy consumption during operation
Compact equipment, modular operation
Not necessary to add other chemicals
Simple process, easy maintenance
Low processing capacity, poor stability, and difficulty in maintaining long-term operating performance
Blockages caused by impurities in the airflow
Low technical maturity
Cryogenic technologyNo chemical reagents required
Excellent CO2 purity
Suitable for high CO2 concentrations
High operating costs
Not suitable for low CO2 levels
Moisture must be removed in advance
Calcium loopingLow raw material prices
Environmentally friendly
Adsorption agent multiple cycles lead to sintering problems, affecting CO2 absorption
Equipment wear issues
Chemical looping combustionLow gas separation costs
No additional energy input required
System energy efficiency
Insufficient stability of oxidation carriers
Not yet industrialized
Equipment wear issues
Table 2. Advantages and disadvantages of different low-energy regeneration methods.
Table 2. Advantages and disadvantages of different low-energy regeneration methods.
MethodAdvantagesDisadvantagesRegeneration Energy Consumption (GJ/t CO2)Reduction in Regeneration Energy Consumption Compared to MEA (%)References
Absorbent Optimization1. Reduces regeneration energy consumption by lowering desorption heat or vaporization heat
2. Fast reaction rate
3. Mature technology
4. Low corrosion
1. Absorption of CO2 increases viscosity, affecting reaction rate.
2. Poor thermal stability.
1.01~2.8525~75%[35,87,126]
Adding catalysts or other materials1. Reduces regeneration energy consumption by lowering desorption heat
2. Fast desorption rate
3. Desorption can occur at lower temperatures
Catalysts face issues such as deactivation and agglomeration.2.28~2.8525~40%[105,127]
Technology enhancement1. Make full use of system waste heat
2. Reduce total energy consumption
1. Need to make certain modifications to existing processes
2. Gas separation issues
3.05~3.605~20%[110,114]
Chemical desorption1. Significantly reduce renewable energy consumption
2. Achieve CO2 sequestration
3. Save equipment costs and operating costs
1. Low mineralization efficiency
2. Insufficient mass and heat transfer
3. Need for constant supply of calcium source
-No regeneration required[118,120,124]
Table 3. Global carbon capture projects in the steel industry [16,25,137].
Table 3. Global carbon capture projects in the steel industry [16,25,137].
Facility NameCountryOperational YearCarbon Capture MethodsCapture Capacity (Mtpa CO2)
ADNOC Al-ReyadahUnited Arab Emirates2016MEA0.8
Baotou SteelChina2025Steel slag carbonization0.5
Nucor Steel DRIUnited States2026_0.8
Indiana Burns Habor CaptureUnited StatesUnder EvaluationSolvents2.8
ArcelorMittal Texas (formerly voestalpine Texas)United StatesUnder EvaluationAir Liquide’s pressure swing adsorption-assisted Cryocap™ technologyUnder Evaluation
ArcelorMittal Sestao CCSSpain2025_Under Evaluation
Japan Malaysia steel CCSJapan, MalaysiaUnder Evaluation_Under Evaluation
Table 4. Cost of carbon capture technology for flue gas in the iron and steel industry [129,132].
Table 4. Cost of carbon capture technology for flue gas in the iron and steel industry [129,132].
Carbon Capture TechnologyMaterialsCO2 ConcentrationCarbon Capture EfficiencyCarbon Capture Cost (USD/t CO2)
Absorption30 wt% MEA24 vol%90%41.92
30 wt% MEA6 vol%90%65.08
50 wt% MDEA-90.24%-
30 wt% MEA4.8 mol%90%73.5
30 wt% MEA27.3 mol%90%55.3
MembranePolyimide membrane4.8 mol%90%271.7
Polyimide membrane27.3 mol%90%41.7
Calcium loopingCalcium-based sorbent-95.44%-
Steel slag--8~104 (No costs for CO2 transportation or storage)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ren, L.; Cheng, S.; Xie, T.; Zhang, Q.; Li, R.; Yue, T.; Cai, C. Low-Energy Regeneration Technologies for Industrial CO2 Capture: Advances, Challenges, and Engineering Applications. Sustainability 2025, 17, 9796. https://doi.org/10.3390/su17219796

AMA Style

Ren L, Cheng S, Xie T, Zhang Q, Li R, Yue T, Cai C. Low-Energy Regeneration Technologies for Industrial CO2 Capture: Advances, Challenges, and Engineering Applications. Sustainability. 2025; 17(21):9796. https://doi.org/10.3390/su17219796

Chicago/Turabian Style

Ren, Le, Sihong Cheng, Tao Xie, Qianxuan Zhang, Rui Li, Tao Yue, and Changqing Cai. 2025. "Low-Energy Regeneration Technologies for Industrial CO2 Capture: Advances, Challenges, and Engineering Applications" Sustainability 17, no. 21: 9796. https://doi.org/10.3390/su17219796

APA Style

Ren, L., Cheng, S., Xie, T., Zhang, Q., Li, R., Yue, T., & Cai, C. (2025). Low-Energy Regeneration Technologies for Industrial CO2 Capture: Advances, Challenges, and Engineering Applications. Sustainability, 17(21), 9796. https://doi.org/10.3390/su17219796

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