Next Article in Journal
Dynamic Simulation of Energy Scenarios in the Transition to Sustainable Mobility in the Ecuadorian Transport Sector
Previous Article in Journal
Solar Self-Consumption and Urban Energy Vulnerability: Case Study in Lisbon
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Comparing CO2 Storage and Utilization: Enhancing Sustainability through Renewable Energy Integration

by
Jose Antonio Garcia
1,2,
Maria Villen-Guzman
1,*,
Jose Miguel Rodriguez-Maroto
1 and
Juan Manuel Paz-Garcia
1
1
Department of Chemical Engineering, University of Malaga, 29071 Malaga, Spain
2
Carbotech Gas Systems GmbH, Kronprinzenstraße 30, 45128 Essen, Germany
*
Author to whom correspondence should be addressed.
Sustainability 2024, 16(15), 6639; https://doi.org/10.3390/su16156639 (registering DOI)
Submission received: 5 June 2024 / Revised: 27 July 2024 / Accepted: 30 July 2024 / Published: 3 August 2024
(This article belongs to the Topic CO2 Capture and Renewable Energy)

Abstract

:
Addressing the environmental challenges posed by CO2 emissions is crucial for mitigating global warming and achieving net-zero emissions by 2050. This study compares CO2 storage (CCS) and utilization (CCU) technologies, highlighting the benefits of integrating captured CO2 into fuel production. This paper focuses on various carbon utilization routes such as Power-to-Gas via the Sabatier reaction, indirect production of DME, and Power-to-Fuel technologies. The maturity of these technologies is evaluated using the Technology Readiness Level (TRL) method, identifying the advancements needed for future implementation. Additionally, global regulations and policies surrounding carbon capture and storage are reviewed to provide context for their current status. The study emphasizes the potential of CCU technologies to reduce future CO2 emissions by converting captured CO2 into valuable fuels and chemicals, thus supporting the transition to a sustainable energy system. The findings indicate that while CCS technologies are more mature, promising CCU technologies can significantly contribute to reducing greenhouse gas emissions if green hydrogen becomes more affordable. This research underscores the importance of further technological development and economic evaluation to enhance the feasibility and adoption of CCU technologies in the pursuit of long-term environmental sustainability.

1. Introduction

1.1. Climate Change and Global Warming

Anthropogenic climate change has become a critical issue, causing detrimental effects on global climate, ecosystems, and species across terrestrial and marine environments [1,2,3]. Climate change includes various aspects such as global warming, shifts in weather patterns, changes in precipitation, more frequent extreme weather events, sea level rise, and alterations in ecosystems and wildlife. Global warming refers specifically to the increase in Earth’s average surface temperature due to the accumulation of greenhouse gases (GHGs) in the atmosphere. This phenomenon is primarily driven by human activities such as burning fossil fuels, deforestation, and industrial processes, leading to a measurable rise in temperatures.
The average global temperature has increased by approximately 1.2 °C above pre-industrial levels, and it is projected to rise further if greenhouse gas emissions are not significantly reduced. The current action plan aims to keep the increase in average global temperature below 1.5 °C [4]. In 2022, CO2 emissions approached 40 billion tons per year (40 × 10 9 t/yr). When considering all GHGs converted to their carbon dioxide equivalent (CO2e), global yearly emissions are estimated to be around 50 billion tons per year (50 × 10 9 t/yr) [5]. The objective is to cut GHGs to 1990 levels by 2030, effectively halving current annual CO2 emissions (to ≈22 × 10 9 t/yr), and achieve net-zero emissions by 2050 [6,7]. However, estimates suggest that 2030 global emissions will far exceed the current goal [8]. Therefore, this ambitious goal necessitates a multifaceted approach, incorporating advancements in renewable energy, carbon capture and storage (CCS), carbon capture and utilization (CCU), and energy efficiency technologies.

1.2. Global Regulatory and Policy Frameworks: Recent COP Outcomes

The Paris Agreement, adopted at the 21st Conference of the Parties (COP21) in December 2015, is a landmark international treaty to combat climate change and accelerate actions towards a sustainable low-carbon future. The Agreement’s central goal was to limit global temperature rise to well below 2 °C above pre-industrial levels, with efforts to cap the increase at 1.5 °C. To achieve this, countries committed to nationally determined contributions, which outline their plans to reduce GHG emissions and adapt to climate impacts. The Paris Agreement also emphasized the importance of climate finance, technology transfer, and capacity building to support developing countries’ climate efforts.
Key outcomes from COP22 to COP25 included the focus on implementing the Paris Agreement, highlighted by the Marrakech Partnership for Global Climate Action and the detailed rulebook established in the Katowice Climate Package. The Talanoa Dialogue and the Gender Action Plan were significant from COP23, promoting inclusive discussions and gender-responsive policies. COP24 emphasized a just transition to low-carbon economies and underscored the urgent need for action following the IPCC’s 1.5 °C report. COP25, while struggling with agreements on carbon markets advanced the Santiago Network for loss and damage and underscored youth engagement in climate action.
COP26 (Glasgow, 2021) reaffirmed the Paris Agreement’s goals. COP26 was the first to explicitly mention coal, oil, gas, and fossil fuels as the primary causes of global warming. The COP26 agreement included the intention to “phase down” rather than shut down coal power plants, with CO2 emissions to be neutralized by CCS technologies despite their high costs [9]. COP27, held in Sharm El Sheikh, Egypt, in autumn 2022, continued these discussions. A week before the conference, the United Nations Environment Program (UNEP) reported that there was no credible pathway to limiting global warming to 1.5 °C, and mitigation efforts since COP26 had been “woefully inadequate” [10]. The agreement aimed to uphold the goal of limiting the temperature increase to 1.5 °C. The COP27 implementation plan emphasized comprehensive strategies incorporating CCS and CCU technologies [11]. COP28, held in Dubai in 2023, further advanced the climate agenda by instituting a more rigorous global review process, setting specific targets for emission reduction and renewable energy expansion, and emphasizing the transition from fossil fuels. Governments recognized the importance of emerging renewable technologies, such as low-emission hydrogen and CCS, in achieving environmental targets [12].

1.3. Captured CO2 Storage and Utilization

CCS is becoming an essential approach to reducing GHG emissions. CCS consists of storing CO2 in a suitable geological sink, which is kept for an extended period. The options are mainly underground: exhausted oil and gas fields, deep coal beds, aquifers, and salt caverns. In some of these cases, CO2 might react chemically with minerals in the rock. Thus, CO2 control over geological time is needed [13,14].
As the world’s largest CO2 emitter, China has set ambitious targets to reach its carbon peak by 2030 and achieve carbon neutrality by 2060. The EU has updated its regulatory framework, including Directive 2009/31/EC on the geological storage of CO2, facilitating CCS in sub-seabed geological formations. This directive enforces stringent permitting, monitoring, reporting, and verification standards to ensure CO2’s permanent isolation in CCS projects.
In the United States, the Bureau of Ocean Energy Management (BOEM) oversees leasing, licensing, and regulation within federal waters on the Outer Continental Shelf, developing a legal framework for offshore CCS facility leases and applying CCS technologies. Japan has introduced regulations under the London Protocol for licensing seabed CO2 storage projects, managed by the Ministry of Economy, Trade, and Industry alongside the Ministry of the Environment. These regulations mandate a comprehensive submission of storage process information and an environmental impact assessment before licensing. In Australia, CCS projects in the Commonwealth marine area are governed by the Offshore Petroleum and Greenhouse Gas Storage Act of 2006, allowing CO2 storage in deep geological formations beyond state jurisdictions [15].
CCS is expected to be the predominant method for large-scale CO2 reduction due to its long-term and secure storage capacity. While CCS is expected to lead in mitigating climate change, CCU is vital, as it reduces emissions and creates economic value from CO2, encouraging broader adoption and innovation in carbon management.
CCU consists of converting the captured CO2 into useful products, such as fuels, chemicals, building materials, and other industrial products. Converting the captured CO2 into valuable chemicals requires H2 production, preferably green H2, produced exclusively with renewable electricity via water electrolysis. Consequently, some CCU technologies present the challenge of producing much hydrogen at a low and competitive price.
Transforming the energy sector is crucial to achieving climate goals. Although the share of electricity produced from renewable sources has increased, it often remains below average demand in most countries. To bridge this gap, flexible carbon or gas power plants match production with demand, primarily due to the difficulty of storing excess electricity generated from renewable sources. This reliance on energy sources such as natural gas led to Europe’s 2021–23 energy crisis [16].
It is relevant to remember that clean energy sources like wind, solar, or hydro-energy depend strongly on weather conditions. Thus, the energy system cannot rely strongly on them. That is why a power system with a high fraction of renewable energy sources may suffer intermittent overproduction, also called a surplus of electricity production (SEP). Sometimes, the SEP may exceed the technical capacity for usage or exportation, creating the critical electricity excess production (CEEP) issue. CEEP represents a low-cost electricity source that has to be instead stored or dissipated [17,18].
A new approach involves significantly increasing energy production from renewable sources to exceed demand, thereby reducing dependency on fossil fuels. The surplus of green electricity, which can vary with weather conditions, would be used to produce green H2. This transition from relying on natural gas for demand surges to utilizing renewable excess represents a sustainable step forward. It aligns energy production with environmental goals, enhances the energy system’s resilience, and reduces fossil fuel dependency. Accordingly, energy storage technologies, classified as Power-to-Power, Power-to-Heat, Power-to-Fuel, and Power-to-Chemicals, also play a crucial role in utilizing captured CO2. These technologies use various energy sources, including electrochemical, chemical, mechanical, and thermal storage [18].
Producing fuels from CO2 and green H2 has two benefits over CCS. First, it reduces CO2 emissions to the atmosphere and transforms the CO2 molecules into renewable fuels. Second, it stores excess renewable electricity at peak production [19]. However, carbon utilization requires, on many occasions, intermediate storage before further transforming CO2 into valuable products. Some processes might require CO2 transportation from the capturing facility to the usage facility, so intermediate CO2 storage is needed before injecting CO2 into the transportation equipment. Furthermore, intermediate CO2 storage can also be applied after the CO2 capture unit to offer a constant flow and condition modification before being introduced into the usage facility. Thus, CCU and CCS can be integrated into carbon capture storage and utilization (CCUS). CCUS includes CO2 capture, CO2 transport, CO2 storage, and final usage. This route requires a considerable amount of energy, which decreases the efficiency of industrial facilities. Many CCUS technologies are expensive and not economically competitive, which caused the main challenge for implementation [20].
Nevertheless, due to the rise of carbon emission tariffs, this situation has changed in recent years, particularly in Europe, through the European Emissions Trading Systems (ETS). The price increased from about EUR 25/t in 2020 to over EUR 80/t in December 2021 [15]. Table 1 shows the carbon prices in different markets per tonne of carbon emissions.
The CCU perspective might provide an opportunity to use CO2 effectively and is a promising way to obtain economic and environmental incentives such as green credits and decrease operative costs, reducing the CO2 emission tariffs and opening secondary business lines in several industries [22].
This review analyzes the usage of captured CO2 generated as a by-product of industrial processes instead of isolating the CO2 in storage for a long time. The pathways and technologies available to capture CO2 are explained in the previous review published by this research team in [23]. The paper reviews the options for storage and utilization of CO2, assuming that the CO2 has been captured. The manuscript provides a comprehensive analysis and comparison of CCS and CCU technologies, emphasizing the potential of integrating renewable energy sources to enhance sustainability. It highlights the benefits of converting captured CO2 into valuable fuels and chemicals. The study includes a detailed assessment of these technologies’ technological maturity, economic feasibility, and environmental impact using the Technology Readiness Level (TRL). The primary objective of the manuscript is to evaluate and advocate for the transition from traditional CO2 storage methods to more innovative CO2 utilization techniques.

2. CO2 Storage

CCS involves the sequestration of CO2 in geological formations, ensuring its long-term isolation from the atmosphere. The primary storage options are underground, including depleted oil and gas reservoirs, deep coal seams, aquifers, and salt caverns. Large amounts of CO2 can be permanently stored in geological formations such as depleted oil or gas fields, unmineable coal beds, aquifers, or saline caves without practical use. Within these geological settings, CO2 may undergo chemical reactions with the surrounding minerals, necessitating meticulous oversight to ensure its stability and containment over geological timescales [13,14].
By 2100, an estimated 2700 Gt of global storage capacity is needed to reach the climate goals [24]. The global capacity for CCS is approximately 40–50 million tons per year (≈0.1% of the current emissions). The IEA underscores the necessity for CO2 storage, growing from today’s capacity to over 5 Gt/yr by mid-century, making it a global industry essential for emission reduction across the energy system. The potential storage capacity globally far exceeds the required amount, with estimates ranging between 8000 and 55,000 Gt, showcasing the vast capability for CO2 sequestration to support deep emission reductions and carbon removal efforts [25].
Nearly 200 new CCS projects were announced during 2022, meaning 60 new CCS facilities since 2021, including projects with transportation and storage facilities with no CO2 capture. These plants’ number and CO2 storage capacity are displayed in Figure 1. The most mature CO2 storage technologies are presented in this section.
When selecting the appropriate geological formation for CCS, several requirements must be considered; rock permeability, thickness, and porosity are critical factors [27]. Also important are storage capacity, the potential for leakage, proximity to the CO2 source, socio-political conditions, and hydrological aspects. Common challenges in CCS implementation include improving efficiency, ensuring long-term safety, navigating government regulations and costs, selecting suitable sites, preventing leakages, and gaining public acceptance. During the carbon storage process, failures in caprock integrity, significant container faults, and valve malfunctions can cause leakages, leading to groundwater pollution. Therefore, advanced sealing and pressure management technologies must be developed to mitigate these risks [28].
While not covered in this review, carbon sequestration in natural sinks, such as the soil of reforested forests, is a crucial strategy for mitigating climate change. Reforestation enhances soil carbon storage capacity, as trees absorb CO2 during photosynthesis and transfer carbon to the soil through their roots and leaf litter. This process not only captures atmospheric CO2 but also improves soil health and biodiversity. Studies estimate that global reforestation could sequester up to 205 Gt of CO2 by 2100. This approach complements technological solutions in achieving carbon neutrality and promoting sustainable land management.

2.1. Enhanced Oil and Gas Recovery

Enhanced Oil Recovery (EOR) and Enhanced Gas Recovery (EGR) involve extracting oil and gas from crude oil fields that cannot be exploited using standard methods. The three primary EOR techniques are gas, chemical, and thermal injection. Gas, chemical agents, or heat is applied to the retained oil or natural gas to facilitate extraction. Currently, the primary use of captured CO2 is for EOR and EGR [23].
Despite utilizing captured CO2 to extract oil and natural gas, these technologies also serve as storage methods. In CO2-EOR, CO2 is injected into a crude oil reservoir to enhance oil extraction. The injected CO2 increases the field pressure, providing the driving force to extract residual oil and natural gas (see Figure 2). CO2 is typically injected in its supercritical phase, which lowers the viscosity of the displaced oil from the rock pores in a single-phase drainage process. The pressure is generally around 75 bar, and the temperature is approximately 70 °C to ensure suitable miscible conditions within the oil field.
CO2-EOR is currently one of the most profitable ways to add value to captured CO2. Economic incentives exist to recover and recycle as much CO2 as possible, thereby reducing the operational costs of the process [29,30]. This makes CO2-EOR a valuable technology for enhancing hydrocarbon recovery and providing a viable means for CO2 sequestration.
CO2-EOR aims to extract crude oil while storing some of the injected CO2 in depleted oil and gas fields. Consequently, EOR does not significantly contribute to the abandonment of fossil fuels and is associated with substantial water usage. Therefore, the net impact on emissions depends on the source of the CO2 and the end use of the recovered oil and gas. It is crucial to evaluate whether this technique results in a net negative emission solution [30,31]. These considerations are essential in transitioning to renewable energy, enhancing CO2 utilization, and fostering a circular carbon economy.
Figure 2. Enhanced Oil Recovery illustration. Adapted from [32].
Figure 2. Enhanced Oil Recovery illustration. Adapted from [32].
Sustainability 16 06639 g002
CO2-EOR has been practiced for several decades, reaching a TRL of 9. The first industrial CO2 injection into a gas field started in 1996 in Norway’s Sleipner Vest gas field. The captured CO2 was injected into saline aquifers near the production gas jacket with more than 20 Mt CO2 injected [15]. There have also been past pilot projects in the US, concretely in the Gulf of Mexico [30]. There are several challenges for offshore CCS-EOR [33]:
  • Higher CAPEX and OPEX compared to the onshore CCS-EOR.
  • Limited CO2 supply. Transport to the offshore facility might increase the overall cost.
  • Limited space for retrofitting new equipment and pipelines. Replacing systems with new technology increases the cost since much more equipment needs to be replaced.

2.2. Enhanced Coal Bed Methane Recovery

CO2-enhanced coal bed methane recovery (CO2-ECBMR) is a technique that involves injecting CO2 into deep coal beds to enhance the recovery of methane from the coal’s porous structure. In this process, CO2 diffuses into the coal’s pore layers and chemically binds to sulfur, demonstrating an affinity almost twice as effective as CH4. The costs associated with CO2 injection and storage can be partially offset by the methane recovered from abandoned coal mines [34].
In conventional coal bed methane recovery (CBMR), methane production efficiency is around 50% due to the rapid reduction in reservoir pressure. However, methane recovery from unconventional CBMR projects is currently only about 6–9%. CO2-ECBMR enhances methane production by lowering the partial pressure of methane in the free gas phase while maintaining overall reservoir pressure [35]. According to Liang et al. [36], more than 60% of high-rank coal gas from closed pores contributes to CH4 and CO2 injection.
CO2-ECBMR has been adopted for many years, with several commercial extraction sites worldwide, primarily in the USA. White et al. highlighted several issues with this practice, such as environmental health and safety, storage integrity, and potential capacity [37]. Unfortunately, the high CO2 absorption capacity can lead to swelling of the coal matrix, which reduces coal permeability and pore size [38].
The estimated current capacity for Enhanced Coal Bed Methane Recovery (ECBMR) in terms of CO2 sequestration is poorly documented. If scaled up to commercial levels, a well-optimized ECBMR project could target CO2 sequestration capacities of 1–2 million tons per year or more, depending on the size and characteristics of the coal bed. Global theoretical storage capacity for CO2 in coal beds could range from hundreds to thousands of Gt of CO2 [35].

2.3. Aquifer Storage

Captured CO2 can be stored in aquifers through injection wells (see Figure 3). Depending on the output pressure of the capture facility, a compression system is needed on-site. The depth range depends on the aquifer’s geological formation. This system consists of an injection well, an injection pump if the pressure needs to be increased, monitoring in the excellent cellar, and different monitoring systems spread out on the surface of the anticipated delineation of the CO2 plume [39].
This method has been used in Europe since 1953 for natural gas storage. Previously, town gas containing 50–60% H2 was stored in aquifers. The company E.ON has operated natural gas storage in Hähnlein, Germany, since 1960, while Dong Energy (now Ørsted) started the operation of natural gas storage in Senlille, Denmark, in 1989. The TRL of this method is 9 [40].
Figure 3. Model of CO2 storage in saline aquifers. Adapted from [41].
Figure 3. Model of CO2 storage in saline aquifers. Adapted from [41].
Sustainability 16 06639 g003
Due to the CO2 injection, the pressure within the reservoir increases. To balance that pressure, a specific volume of water can be extracted and brought to the surface through a separate production well. This produced water could potentially be utilized in a geothermal system, thereby enhancing the economic feasibility of the overall system. However, pressure management is complex due to the lack of regulatory guidelines on permissible pressure increases. Also, managing pressure involves handling large volumes of water at the surface, presenting logistical and environmental challenges [39].
Injecting CO2 into a saline aquifer can impact the petrophysical properties of the reservoir rocks through complex geochemical reactions. The CO2 dissolves in the brine and subsequently precipitates as carbonate minerals, partially deleting the stored CO2. Continuous CO2 injection and the resultant lower brine pH increase the risk of corrosion. Furthermore, carbonate precipitation reduces the porosity of the saline aquifer, thereby diminishing the system’s energy storage capacity [42].
Once the CO2 is stored in the aquifers, it can be extracted for further use. However, pumping the CO2 out of the aquifers for further utilization produces moderate results. Some studies demonstrate that 33–50% can be pumped up, but a CO2 recovery of 50% is required to obtain favorable conditions. Additionally, part of the CO2 is lost due to several factors. For example, CO2 can react with minerals and be mineralized. Also, if the temperature drops below 10 °C, the CO2 can form hydrates with water, and the CO2 can be dissolved in the formation water [43,44].

2.4. Salt Cavern Storage

Since the 1970s, natural gas has been stored in salt caverns in the USA and Europe, with more than 300 such caverns utilized in Europe alone. Town gas and hydrogen (H2) have also been successfully stored in salt caverns under similar operational conditions, albeit with different material and safety measures at the surface [40]. Due to its maturity and proven reliability, this technology has achieved a Technology Readiness Level (TRL) of 9.
Salt caverns are artificial cavities in salt deposits, typically found in salt domes. The process involves drilling a well into the salt dome and injecting water to dissolve the salt, thereby creating a cavity with resilient walls that resist reservoir degradation [45]. While these saline caverns have no other commercial value, they can be repurposed for CO2 storage. However, storing CO2 in salt caverns has historically been more expensive than other technologies such as aquifers and EOR [29,46,47]. Despite the higher costs, CO2 recovery for utilization after storage is feasible. Various trapping mechanisms exist when CO2 is injected into a saline cavern, providing different levels of security and efficiency for long-term storage; see Table 2.
Estimates suggest that the global theoretical capacity for CO2 storage in salt caverns could be tens to hundreds of Gt. However, at the moment, the actual amount of CO2 being stored in salt caverns is relatively small compared to their theoretical capacity. Several pilot and commercial projects have been established for CO2 storage in salt caverns. The Statoil Sleipner project, for instance, stores about 1 Mt CO2/yr in a 1 km deep cavern in the Utsira Sand formation in the North Sea, with a storage capacity of 660 million cubic meters [51,52,53]. Another notable example is the Latrobe Valley commercial project, which started in 2015 in Victoria, Australia, with a capacity of 13 Mt CO2/yr [54].
In Lille Torup, Denmark, seven salt cavern storage sites are located at depths between 1000 and 1700 m. These caverns are 300–350 m high and have 40–60 m diameters, with an estimated total capacity of about 0.3 Mt of supercritical CO2 per cavern [55]. The caverns are naturally tight due to the unique properties of rock salt, which can be structured in salt diapirs, bedded salt, and salt domes. Optimal pressure for storage depends on the cavern depth and the type of fluid or gas being stored, commonly ranging from 65 to 180 bar at depths of around 1000 m [40]. It is recommended to store CO2 in the dense liquid or supercritical phase [47].
One of the main drawbacks of this technology is the large amount of wastewater produced. In particular, during the creation of the cavern, significant quantities of brine are pumped to the surface. Proper handling of the brine is crucial to avoid environmental impact. This is typically achieved by mixing the brine with seawater before discharge into the sea [40,46,47].

2.5. Deep Ocean Storage

Oceans, covering more than 70% of the Earth’s surface, are the largest natural carbon sinks. There are two primary processes for ocean CO2 storage: (1) ocean fertilization and (2) direct CO2 injection [56].
Ocean fertilization is a biological CO2 storage method that promotes the growth of oceanic organisms, such as phytoplankton, which integrate CO2 into their metabolisms through photosynthesis. By stimulating the growth of these organisms, more CO2 is absorbed from the atmosphere and sequestered in the ocean’s biological systems. This process enhances the natural biological pump, which transports carbon from the surface ocean to the deep ocean, where it can be stored for centuries or longer.
In the direct CO2 injection method, captured CO2 is transported to the ocean via pipeline or ship. CO2 can be injected in various forms, including gas, liquid, solid, or supercritical fluid, as shown in Figure 4. CO2 can also be injected into gas hydrates, crystalline solids formed from water and gas. Although promising, this technology is not yet ready for widespread practical application [56]. At depths greater than 3 km, CO2 becomes liquefied and is sequestered at the ocean bed due to its higher density compared to seawater [47,57].
Sheps et al. [58] proposed criteria to compare CO2 storage systems, highlighting that ocean CO2 sequestration offers large capacity and medium permanence, with a 30–500-year time scale. It is considered an irreversible process at a relatively low cost. However, this approach is more controversial than other storage methods due to potential environmental impacts.
At depths greater than 3 km, CO2 becomes denser than seawater, facilitating its sequestration at the ocean bed. The theoretical maximum capacity for CO2 storage in the deep ocean is vast, estimated to be thousands of Gt. One major concern is the impact on seawater chemistry. Injected CO2 reacts with seawater to form carbonic acid (H2CO3) and hydrogen ions (H+), leading to ocean acidification. This acidification reduces the seawater pH and can significantly harm marine ecosystems, affecting species such as coral, shellfish, and plankton sensitive to pH changes [56,59,60]. Although the vast volume of seawater could theoretically buffer these changes over time, the localized and long-term impacts on marine life could be severe, with potential consequences for the broader marine food web and biodiversity [56].

3. CO2 Utilization Technologies

As previously discussed, the CCS approach alone does not solve the global warming problem. Even technologies like EOR fail to help us reach the net-zero CO2 emissions target, as they lead to the production of more fossil fuels. Moreover, techniques like deep ocean storage can harm natural media. Additionally, CCS has geological constraints since there is not enough for permanent storage at current CO2 emission rates, as illustrated in Figure 5. This means that we need to look at alternatives to CO2 storage.
CCU is an innovative approach that focuses on repurposing captured CO2 for various applications rather than storing it. The objective is to purify and utilize CO2 in industrial processes or to transform it into valuable products such as chemicals, materials, or fuels. Capturing CO2, converting it into fuels, and subsequently using these fuels delays the ultimate release of CO2 into the atmosphere.
This approach primarily reduces the consumption of fossil fuels, thereby decreasing overall CO2 emissions. It promotes the concept of an artificial CO2 cycle, which involves capturing CO2, converting it into valuable products, optionally storing it temporarily, using it, and then capturing it again. While this cyclic process helps mitigate CO2 emissions, it is essential to recognize that it does not eliminate them. Consequently, CCU should be integrated with other strategies to achieve comprehensive emission reduction targets.

3.1. Non-Energy Applications

The global consumption of CO2 as a chemical reagent has risen in recent years. It was about ≈150 Mt/yr by 2000 and increased to ≈250 Mt/yr by 2022. In 2015, the major markets were urea for fertilizer production and EOR; see Figure 6a. Based on an average year-on-year growth rate of 1.7%, global CO2 consumption is expected to rise in the years to come, as shown in Figure 6b. The three major CO2 markets are the USA (33%), China (22%), and Europe (16%) [62,63].
The CCU approach is expected to produce revenues, reducing pressure on energy costs, and it can create economic benefits and job opportunities [64]. The Global CO2 Initiative shows an annual market on CO2-based products of USD 0.8–1.1 trillion using about 10% of the global CO2 emissions. Table 3 displays how CO2 is already used in various sectors, such as oil, power, chemical, food, steel, and pharmaceutical industries.
The highest potential for direct CO2 utilization lies within the oil and chemical industries. Urea production represents a significant demand area, estimated at up to 300 Mt/yr. Polymer processing and synthesizing renewable fuels, such as methanol, are key fields for CO2 utilization. The cement industry also demonstrates substantial potential, with an estimated demand of approximately 300 Mt/yr. The food sector has moderate potential for CO2 utilization, primarily in applications such as packaging, decaffeination, horticulture, and beverage carbonation [65,66].
Table 3. State of the art for CCU [67,68,69,70,71,72].
Table 3. State of the art for CCU [67,68,69,70,71,72].
PotentialProcess/ProductSector
HighOil extractionEnhanced Oil, Gas and Coal Recovery
Stimulation/Fracturing of oil and gas
Polymer Processing
Chemicals and Fuels (Methanol, methane, CO, fertilizers and derivatives)
MediumFood processingCoffee decaffeinating
Wine production
Beverage carbonation
Dry ice production
Horticulture (greenhouses)
Food packaging and preservation
MediumMineralizationCalcium and magnesium carbonate for use in cement
Calcium bicarbonate
Bauxite residue treatment
CO2 concrete curing
MediumPowerWorking medium in CO2 cycles
Heat pumps
LowSteelBottom stirring agent in basic oxygen furnaces
Injection to metal casting
Hardening sand cores and molds
Chilling medium
LowPharmaceuticalChemical synthesis
Supercritical Fluid Extraction
Product transportation
Inerting
LowPulp and PaperpH reduction during washing
LowEnergy cropsAlgae cultivation
LowOtherElectronics (printed circuit manufacture)
Pneumatic (working medium in hand tools and equipment)
Fire extinguishers, fire suspension
Urea production
Flavors, Fragrances
Blanket Products
Aerosols and propellants
Soda ash production for glass industry
Welding (shield gas)
Dry gas cleaning
Refrigerant gas
Water treatment

3.2. CO2 Applications in the Energy Sector

The US Department of Energy actively supports research into CCU and reuse technologies [73]. Similarly, the European Commission has launched several initiatives under the Horizon 2020 framework, including projects such as BIOTEC, LCE 25, SPIRE, and NMBP 19 and 20, to explore CCU pathways for the production of chemicals, intermediates, and alternative fuels [74].
There is significant potential for CCU in the power and fuel sectors, with applications spanning coal, biomass, natural gas, waste incineration, and small power plants, as well as nuclear plants [64,65,73,75,76].
As previously mentioned, expanding renewable technologies is crucial for reducing greenhouse gas (GHG) emissions. However, the expansion of renewable energy sources intensifies the fluctuation of supply and demand curves due to the unpredictability of weather-dependent energy production. Power systems with a high fraction of energy produced by renewable sources, such as wind or solar, may experience periodic surpluses in electricity production (SEP). In some cases, SEP exceeds the capacity of existing transmission lines, making it impossible to export the surplus. This unexportable SEP is termed critical electricity excess production (CEEP) and represents low-cost electricity that must either be stored or dissipated [17].
The term “green H2” refers to hydrogen produced via water electrolysis using renewable energy, preferably from CEEP. Green H2 is recognized as a versatile energy carrier to address decarbonization and support the renewable energy transition. H2 can be applied in the chemical sector and the steel industry and can also be used as a direct fuel for transportation or electricity production via fuel cells [77]. In addition to its use as a fuel, H2 can be combined with CO2 (preferably captured CO2) to produce a wide range of energy carriers through fuel synthesis processes suitable for both mobile and stationary applications [78]. Currently, CO2 hydrogenation enables the production of chemicals such as methane, methanol, ethanol, and formic acid, which can serve as alternatives to fossil fuels or feedstocks for producing complex chemical compounds.
The group of technologies that transform electrical energy into valuable chemicals using H2 is known as “Power-to-X”, where X represents various chemical products, gases (e.g., methane), liquid fuels, or electricity (see Figure 7). The Power-to-Gas approach involves using green H2 to produce CH4. Although CH4 is a carbon-based fuel, it offers several advantages over H2. For instance, the storage and distribution of H2 present significant technical challenges, whereas there is an existing infrastructure for methane transportation and distribution via pipelines, liquefied natural gas (LNG) carriers, and gas turbines for converting methane into heat and electricity. The produced Synthetic Natural Gas (SNG) can be injected into the natural gas network [79,80,81].

3.2.1. The Catalytic Sabatier Reaction

The Sabatier process combines CO2 with H2 to produce CH4:
4 H 2 + CO 2 CH 4 + 2 H 2 O
Δ H 298 K = 165 kJ / mol
This reaction is a linear combination of the reverse direction of the steam reforming, Equation (2), and the water–gas shift (WSG), Equation (3), reactions [82]:
CO + 3 H 2 CH 4 + H 2 O
Δ H o = 206.2 kJ / mol
CO 2 + H 2 CO + H 2 O
Δ H 298 K = 41 kJ / mol
The French chemist Paul Sabatier and his co-worker Abbé Senderens discovered this catalytic methanation process at the beginning of the 20th century. The technology has been mainly applied to remove carbon oxide traces from feed gas for ammonia synthesis [83]. Nowadays, the conversion of CO2 to CH4 is gaining interest due to its application in areas such as biogas upgrading, to transform the bio-CO2 into additional biomethane [84]. The Sabatier synthesis has a TRL of 9 [84].
The catalytic Sabatier process is well understood and used for industrial applications such as ammonia production. As the reaction is exothermic, higher temperatures will thermodynamically hinder the process. Therefore, it is necessary to increase the pressure to ensure conversion. Typical operating conditions are between 300 and 800 °C with a pressure of up to 20 bar [85]. The process requires a pretreatment to remove impurities such as H2S and siloxanes [86].
One of the challenges for the methanation process is catalyst development [80]. Several metals may be used as catalysts for the methanation reaction: ruthenium (Ru), rhodium (Rh), cobalt (Co), and nickel (Ni). However, Ni is considered the most optimum catalyst choice, owing to the good CH4 selectivity, high activity, and low price [87].
The highest CH4 yield is obtained at high pressure and low temperature. A yield of 98% is obtained with a temperature below 300 °C and at 10 bar. The yield decreases drastically at temperatures above 600 °C, depending on the pressure [88]. Nevertheless, a high-pressure operation is not economical, and a low-temperature operation requires a highly active catalyst. The methane yield at 99.5% is possible with an energy efficiency of around 77%, as seen in Figure 8:
The excess heat can be used in other parts of the process to improve energy efficiency. Currently, there are different configurations for the catalyst reactors. Generally, fixed-bed reactors present the advantages of easy utilization; thus, it is the most frequently used reactor [85]. To optimize the catalyst behavior and heat management, there are different methanation processes based on several fixed-bed reactors in series with intermediate cooling [89]. More recent research has used four fixed-bed reactors in series with integrated air cooling systems with intermediate heaters to pre-heat the gas mixture [90]. New reactors are being developed, such as microreactors and three-phase methanation [80,81,91].

3.2.2. The Biological Sabatier Reaction

This process presents the advantages of minimal chemical applications and lower energy requirements. Despite being known for many years, it is still at the demonstration level [92]. It can be accomplished using two main processes: an in situ reactor and a separate reactor. The in situ process is typically used in anaerobic digestion systems to increase the proportion of CH4 over CO2 of biogas by injecting H2 directly into the digester. This process does not need a second reactor and reduces the investment, but total CO2 conversion is very difficult [87]. Alternatively, a separate biological reactor can convert captured CO2 from biogas or any other source. This option presents the advantage of allowing the conditions to be adjusted according to the requirements of the hydrogenotrophic methanogens. However, the main challenge for this reactor design is the H2 supply for the microorganisms because of the limit rate.
Biological methanation is a simple way to upgrade biogas but harms the anaerobic process due to some inhibition mechanisms. For example, CO2 removal may increase the pH value. Thus, in situ, biological methanation may require co-digestion with acidic substrates or pH control to keep the pH within appropriate levels [93]. The heat excess from the biological methanation is not high enough to overheat the anaerobic digestion process but is still high enough to be worth recuperation for other uses [81].
Biological methanation operates at temperatures of 30–60 °C and atmospheric pressure and, in contrast to the catalytic process, has a high tolerance against pollutants in the feed gas. Previous studies [94,95] showed mesophilic and thermophilic anaerobic digestion to convert CO2 to CH4 by adding H2. At thermophilic temperature (55 °C), bio-conversion of CO2 and H2 produced results 60% higher than at mesophilic temperature (37 °C), providing higher efficiency for biogas upgrading. The CH4 produced can reach around 90–95% of CH4 [93], the rest being a mixture of H2 and CO2. Due to the exothermicity of the biological reaction, the overall energy efficiency is around 80% [86].
Despite biological methanation having low kinetics and flexibility, its use is increasing due to the rise of biogas production [80,81,91]. Consequently, it is possible to conclude that a biological process is best suited for small to medium-sized plants, preferably biogas production plants. In contrast, catalytic processes seem more suitable for larger plant sizes. The fixed-bed reactor is a well-known and well-suited technology for large-scale methanation plants (>100 MW) [87].

3.2.3. Dry CO2 Reforming

Dry CO2 reforming is an attractive solution to produce syngas from CO2 and CH4. The produced syngas can be further used to produce fuels and chemicals via fossil-free petrochemical methods.
CH 4 + CO 2 2 CO + 2 H 2
Δ H 298 K = 247 kJ / mol
The reaction is endothermic and requires a catalyst, typically Ni-based. The catalyst may suffer from fouling due to carbon deposition from undesired reactions such as:
CH 4 C + 2 H 2
This fouling can be avoided by the interaction of a wide size distribution of Ni with the metal oxide surface as a support material [96,97,98]. The selection of MgAl2O4 as a metal oxide can slow the formation of carbon deposits [99,100,101,102]. However, higher catalytic activity has been acquired with the usage of rhodium (Rh), platinum (Pt), iridium (Ir), and lead (Pb) [103]. This reaction can achieve a conversion efficiency of up to 98% at 600–800 °C and 1 atm [30].
The redox cycle with CeO2 offers an attractive thermodynamic conversion over a metal oxide catalyst. In this cycle, the metal oxide takes oxygen from the CO2, forming CO. Secondly, the catalyst is regenerated by releasing oxygen. The CeO2 catalyst forms H2 from water at 900–1000 °C and offers an efficiency of 19–22%. Even though CeO2 is the most commonly studied catalyst for this reaction, this technology has not yet been applied commercially and is still not sufficiently mature [104,105,106].

3.2.4. Methanol Production

Methanol is a suitable energy carrier for the fuel and chemical industry due to its flexible properties [81,107,108]. The main route for methanol production is steam reforming of natural gas, but it can also be produced from non-fossil H2 by water electrolysis, biomass conversion, or solar conversion [109,110,111].
Methanol is also the base to produce many other chemicals. It can be the basis for transportation fuel production, such as polyoxy dimethyl ether, dimethyl ether, and higher-alcohol-content products like butanol [112,113]. It can also be applied as a gasoline additive [86,109,114]. Since methanol has a high octane rating, it can also be used as a gasoline substitute in car engines [113].
Methanol can be produced from syngas. This synthesis takes place at 250–300 °C and 50–100 bar, with the use of Cu/ZnO as catalyst together with Al2O3 as promoter [107,108,109,110,111,115].
CO + 2 H 2 CH 3 OH
Δ H o = 90.5 kJ / mol
Methanol can also be synthesized from CO2 by including an intermediate step of converting CO2 to CO by the reverse water–gas shift (RWGS) reaction, resulting in:
CO 2 + 3 H 2 CH 3 OH + H 2 O
Δ H o = 49.5 kJ / mol
The CO2 hydrogenation reaction produces water as a by-product; thus, a third of the H2 is converted to H2O. This water conversion ratio is much higher than in commercial methanol synthesis from syngas. Furthermore, the thermodynamic properties of direct methanol production from CO2 are not as favorable as those from CO. The methanol yield from CO2 at 200 °C is about 40% less, while the yield from CO is higher than 80%. This is mainly due to the Cu/ZnO catalyst and Al2O, which makes the methanol production from syngas easy, showing lower reactivity from CO2 at temperatures below 250 °C.
Despite the reactivity increase by raising the temperature, this leads to the formation of CO and H2O by RWGS. Consequently, additional H2 is consumed, and the methanol production decreases [109]. Therefore, the efficiency depends enormously on the temperature and can be between 30 and 92%, while higher temperatures drop the efficiency rate. The excess heat can be used in high-temperature electrolysis, improving the overall efficiency of the system [116]. The efficiency for methanol production by direct hydrogenation of CO2 is reported to be around 79% with a methanol selectivity above 99.8%; see Figure 9.
Methanol synthesis from CO2 has become technically competitive with the syngas route due to the development of practical reactors and effective catalysts [22,109,111]. Several companies offer commercial solutions, such as Luigi, Mitsubishi, and Haldor Topsoe [109]. Iceland’s George Olah Renewable Methanol Plant, run by Carbon Recycling International (CRI), is one of the largest commercial production plants, with 4000 tons/y. The facility reacts H2 produced from water electrolysis using electricity from a geothermal power plant and CO2 from an integrated capture system [78,107,110,117]. The TRL of this technology is estimated to be eight and could soon reach nine [108].

3.2.5. Fischer–Tropsch Process

The Fischer–Tropsch process consists of the hydrogenation of CO to produce a variety of higher hydrocarbons (C5+), mainly alkanes and alkenes. The mixture is called syncrude and can be refined into fuels such as diesel or gasoline [114,118,119]. These fuels can be used in internal combustion engines, releasing fewer emissions compared to conventional gasoline or diesel, as Fischer–Tropsch fuels have a lower concentration of N2, aromatics, and sulfur [120,121].
The CO applied for Fischer–Tropsch fuel production can be obtained by various pathways, such as CO2 SOEC electrolysis, biomass gasification producing bio-syngas, and from RWGS reaction from CO2 [114,116,118,120,122]. The resulting syngas is fed to the synthesis reactor to produce a mixture of long-chain hydrocarbons; see Reaction 8:
CO + 2 H 2 CH 2 + H 2 O
Δ H o = 158.5 kJ / mol
This reaction is a simplified version of several simultaneous reactions. The -CH2- is the carbon chain from which higher hydrocarbons are generated (C1–C70+) [123]. Gasoline consists of a hydrocarbon chain between C4 and C12, while the diesel carbon chain is in the range of C10–C20 [124]. The mixture and the carbon chain length depend on operation conditions, catalyst, and the syngas composition [81,118,119,120]. This process can produce other fuels, like aviation biofuels, which have gained relevant interest due to the possibility of reducing emissions in the aviation sector. The synthetic diesel produced by this technology can also be applied in the shipping sector. Thus, all transportation sectors can reduce GHG emissions by integrating this approach.
The Fischer–Tropsch process can operate at different temperatures. The low- temperature Fischer–Tropsch process occurs between 210 and 260 °C, while the high-temperature Fischer–Tropsch process operates at 310–340 °C. The operation temperature determines the fuel properties and the carbon number. High-temperature Fischer–Tropsch fuels are mainly olefins and aromatics, whereas low-temperature Fischer–Tropsch fuels consist primarily of paraffin. Thus, the high-temperature Fischer-Tropsch process is more suitable for gasoline production due to the high octane number of aromatics. The low-temperature Fischer–Tropsch process is applied for diesel production because of the high cetane number of paraffinic compounds [125,126]. Typically, this process applies fixed- or fluidized-bed reactors operating at 10–60-bar pressures with iron-based or cobalt-based catalysts together with a supporter, usually Al2O3 [120,127].
The total efficiency of the Fischer–Tropsch process to produce hydrocarbon with a carbon chain of C5+ is typically in the range of 60–90% with a selectivity of about 85% [81,118,128]. The Low Fischer–Tropsch process selectivity for diesel production is much lower, about 26%; see Figure 10. This result includes waxes and by-products.
The Fischer–Tropsch product is then separated and refined to obtain the final fuel. The excess heat can be used elsewhere in the system to improve efficiency. The surplus heat is estimated to be 6.57 MJ steam/kg CO2. These data are in the range of the heat demand of 5.4–7.2 MJ/kg CO2 at 95 °C for CO2 capture from air. The excess heat can also be applied to the electrolysis process, increasing the final efficiency [112].
The Fischer–Tropsch process or Power-to-X was developed by Franz Fischer and Hans Tropsch in the Kaiser Willhelm Institute for Coal Research at Mülheim (Ruhr) in Germany in the 1920s from hydrocarbon production from coal gasification products [130,131]. Germany applied this process during the Second World War to extend the Nazi’s fuel. This approach has obtained considerable interest in producing sustainable synthetic liquid fuels [132]. Sunfire inaugurated a demonstration plant in Dresden (Germany) in 2014. The facility produces synthetic diesel using the Fischer–Tropsch process. The excess heat of the process is applied to heat the Solid Oxide Electrolysis Cell (SOEC) to produce H2, improving the overall efficiency [116]. This process is entirely commercial, and a variety of fuels are made all over the world. Nevertheless, most facilities use non-renewable feedstock with CO2 from emitting processes. The Fischer–Tropsch process integrated with renewable CO2 pathways is not very mature yet. That is why the TRL is between 5 and 9 [133].

3.2.6. CO2 Electrochemical Reduction

CO2 electrochemical reduction (CO2ECR) can generate valuable products such as CO, methanol, ethanol, formic acid, oxalic acid, methane, and formaldehyde based on a specific reaction promoted by catalysts and solvents in an electrolysis cell. This reaction’s selectivity and faradaic efficiency depend on several parameters: pH, electrode potential, selected electrolyte, and catalyst. The generation of H2 from water is one of the main challenges of CO2ECR due to the energy cost. Nevertheless, this can be avoided using aprotic or non-aqueous solvents and a cathode with minimal H2 over-potential. However, more reliable, scalable, and cost-effective electrolyzer designs are still required for the future application of CO2ECR [134]. Kibria et al. [135] disclose a Life Cycle Assessment (LCA) comparing CO2ECR for fuel and feedstock production. They conclude that CO2ECR must be combined with low-carbon electricity production for lower GHG emission benefits.
CO2ECR reactions occur at the electrolysis cathode surface where CO2 is reduced to form products like HCOOH, CO, or OH. In contrast, oxidation reactions such as oxygen evolution, amine oxidation reaction, and H2 oxidation reaction occur at the anode depending on the electrocatalyst and the anode electrolyte [136,137].
The CO2 molecules are adsorbed to the cathode to an active anion radical form at 25 °C and 1 atm pressure. The electrolyzer’s ion exchange membrane acts as a barrier between the cathode and anode, allowing the transport of selective ions while reducing the transport of all other products, avoiding the reverse oxidation of the product to the anode. An electrolyte solution, such as KOH or NAHCO3, conducts electricity, delivering and dissolving CO2 to the cathode. On the other hand, electrons are transferred by a power source with enough voltage from the anode to the cathode [134]. The CO2ECR includes several steps:
(1)
CO2 transport from gas phase to electrolyte bulk.
(2)
CO2 diffusing through the electrolyte to the electrolyte/cathode interface.
(3)
CO2 adsorption at the cathode.
(4)
Intermediate product production at the cathode, such as CHO, COH, COOH, and CO.
(5)
Intermediate electrons coming from the cathode catalyst.
(6)
Product desorption from the electrode.
(7)
Reduced products transfer into liquid phases or bulk gas.
The electrolysis design controls the amount of electrolyte required for the reaction. The cell design affects the electrolyte requirement, gas diffusion, cathode area, and other parameters to assure the reaction performance at a low cost [30]. The ion exchange membrane, type of catalyst, electrodes, and electrolyzer cell configuration influence the CO2ECR process, including CO2 conversion rate, long-term stability, Faraday efficiency, and energy efficiency. CO2ECR allows a modular design in combination with CO2 capture and CO2 utilization technologies since it operates at mild conditions. To reduce CAPEX in the entire process from CO2 capture to CCU, the CO2ECR process should be operated at a high Faraday efficiency, CO2 conversion rate, current density, and energy efficiency [134]. Roh et al. [138] analyzed the potential of ten products from electrochemical CO2 reduction at TRL 2.

3.2.7. Photocatalytic CO2 Reduction

Photocatalysis is an advanced oxidation process that converts inorganic and organic pollutants into non-hazardous compounds. This method utilizes photon irradiation to accelerate redox reactions, effectively transforming solar energy into chemical energy and lowering the activation energy required for chemical transformations. It is recognized as an advanced, environmentally friendly, and cost-effective technology for reducing pollutants in wastewater treatment applications [139]. CO2 photocatalytic reduction (CO2PCR) involves using a semiconductor catalyst to convert CO2 into valuable products. In photocatalysis, photons of light with energies equal to or greater than the semiconductor’s band gap energy ( E b g ) excite electrons (e) and create holes ( h + ) in the semiconductor’s conduction and valence bands, respectively [30]. CO2PCR is noted for its energy efficiency compared to CO2 electrochemical reduction (CO2ECR), as it operates under milder conditions.
CO2 is a very inert and stable molecule due to its high C-O bond energy (750 kJ/mol), necessitating elevated pressures and temperatures for thermocatalytic conversion. However, CO2 photocatalytic reduction (CO2PCR) can be optimized using semiconductor materials to produce valuable products such as CH4, CH2O, HCOOH, or methanol [140]. Additionally, the CO2PCR process can generate C2 products from C-C coupling reactions, including CH3CH2OH, C2H4, and C2H6 [141,142,143].
Current research focuses on enhancing the energy efficiency of the CO2PCR process on semiconductors. TiO2-based semiconductors are the most extensively studied due to their high photoactivity. Other semiconductor materials, such as CuO and Cu2O, have also garnered significant attention. Further studies have explored materials like Ta2O5, MgO, Ga2O3, In2O3, Bi2MoO6, and Bi4Ti3O12. These materials can be doped with metals such as Fe, Cu, Pd, Cr, Pt, Co, Ni, Ag, or Ru to reduce charge recombination. Additionally, several combined semiconductor systems have been extensively studied to improve process performance in recent decades. These include TiO2/polydopamine, CeO2/3D g-C3N4, CsPbBr3/Bi2WO6, and MXene/Bi2WO6 [139].
The design of a hybrid system based on photoelectrochemical cells combined with a photovoltaic cell has shown promising results by separating the oxidation reduction reactions into two subsystems, thereby increasing process optimization. Additionally, this hybrid approach maintains high solar energy conversion efficiency into fuel without requiring precious materials. Despite these advancements, the photocatalytic reduction of CO2 currently has a low Technology Readiness Level (TRL) of 3–4 [139,144].
The first step in the CO2 photocatalytic reduction (CO2PCR) process is to capture and absorb light to generate efficient photogenerated electrons (e) and holes (h+). This photogenerated e must possess an adequate reduction potential to achieve the desired product selectivity and initiate the specific reduction reactions. However, the process still faces several challenges, including insufficient light absorption, photocorrosion, and low reaction kinetics. The CO2PCR process begins with the adsorption of CO2 onto the photocatalyst’s surface, followed by hydrogenation, deoxygenation reactions, and the formation of C-O bonds with the photocatalyst’s surface atoms. The sequence of hydrogenation and deoxygenation reactions is crucial in determining product selectivity. The adsorption of CO2 can follow three pathways, which dictate the reaction route:
(a)
Carbon coordination.
(b)
O2 coordination.
(c)
Mixed coordination.
Therefore, understanding the fundamental chemical pathways is essential for significantly improving the CO2 photocatalytic reduction (CO2PCR) process. This understanding offers two potential advantages: (1) lowering the key thermodynamic or kinetic barriers of the intermediate reactions to increase the CO2 conversion rate, and (2) regulating the process to generate the desired product. Potential intermediates can significantly impact the reaction pathways, influencing product selectivity [139,143]. Two proposed pathways are prominent in the CO2PCR process: the formaldehyde and carbene pathways. The formaldehyde pathway reduces CO2 to produce CH4, while the carbene pathway can reduce CO2 to produce either CH4 or methanol, depending on the specific conditions. The formaldehyde pathway follows the reaction outlined in Equation (9)
CO 2 HCOOH H 2 CO CH 3 OH CH 4 ,
while the carbene pathways follows Reaction (10):
CO 2 CO C CH 3 CH 3 OH / CH 4 .
Methanol is formed faster than methane with aqueous photocatalysis [145]. This is due to the formation of bicarbonate, carbonic acid, and carbonate species, resulting in the generation of HCOOH and final reduction to CH3OH. The methane production may also produce heavier hydrocarbons like CH3COCH3 and CH3COOH [146].

3.2.8. CO2 Fuel Cells

Molten carbonate fuel cells (MCFCs) can capture and convert CO2 into heat and power [147]. Thus, CO2 fuel cells can also be considered a CO2 capture technology. These fuel cells require H2 at the anode and use CO2 and O2 as input at the cathode. At the cathode catalyst (Equations (12)), CO2 reacts with O2 generating CO 3 2 that migrates through the membrane towards the anode, as shown in Equation (12), and reacts with H2 to produce CO2 and H2O [148]. Consequently, this method allows the simultaneous purification of CO2 containing currents and the production of electrical energy.
A n o d e : CO 3 2 + H 2 CO 2 + H 2 O + 2 e
C a t h o d e : 2 CO 2 + O 2 + 2 e 2 CO 3 2
CO2 in flue gases from power plants or industries can be directly fed into an MCFC, as illustrated in Figure 11. The MCFC electrolyte plays the role of a membrane that allows capturing CO2 from the flue gas with additional heat and power generation. The anodic outflow of CO2 and H2O can be separated by water condensation. MCFC process performance depends on the CO2 concentration. Therefore, as standard flue gases present a relatively low CO2 concentration, between 4 and 15%, CO2 recirculation may be used. Removal of flue gas impurities is required to avoid undesired electrode processes.
The generated heat and power can also be recovered and utilized in a power plant or other applications [148]. A combined cycle power plant integrated with molten carbonate fuel cells (MCFCs) can capture up to 90% of the CO2 in the flue gas while generating additional heat and power, thereby increasing the overall efficiency of the process. This technology is considered to have a Technology Readiness Level (TRL) of 5 [149].

3.2.9. Dimethyl Ether and Dimethyl Carbonate Synthesis

Dimethyl ether (DME) is an organic compound primarily used as an aerosol propellant and a reagent in producing dimethyl sulfate and acetic acid compounds. There are two routes for its production: the direct route (TRL 4), which converts syngas using a bi-functional catalyst (Cu/ZnO/Al2O3), and the indirect route (TRL 9), which involves the dehydration of methanol. Although the indirect route is more commonly employed, the direct route is considered more economically and thermodynamically feasible [81,108]. DME serves as an alternative fuel to conventional diesel. Its combustion produces no aromatics or sulfur, resulting in lower emissions of NOx, SOx, CO, and particulate matter compared to traditional diesel [107,150,151,152]. Additionally, DME can be utilized as liquified petroleum gas (LPG) for heating purposes [150,153].
Dimethyl carbonate (DMC) is a valuable intermediate in organic synthesis and a fuel additive. Its diverse applications stem from its favorable chemical and physical properties [154,155]. DMC can be used in green adhesives and as a gasoline additive, owing to its high oxygen content compared to other chemicals [156,157]. Other applications include the production of antibacterial medicines, veterinary drugs, and insecticides [158,159,160]. One of the most critical future applications for DMC in the energy sector is its use as a battery electrolyte, which can enhance conductivity, current density, and battery life [161]. There are several methods for the synthesis of DMC: the phosgene method, oxidative carbonylation of methanol, transesterification method, and direct DMC production from methanol [162,163,164]. The direct route faces two main challenges. First, the reaction’s inherent limitations and the high inertia of CO2 necessitate the use of an effective catalyst. Second, the efficiency and reusability of the selected catalyst are critical factors that influence the overall feasibility and cost-effectiveness of the process [165]. The direct DMC production (TRL 3) is explained in Equation (13):
CO 2 + CH 3 OH DMC + H 2 O

4. Discussion

The main target set by the Intergovernmental Panel on Climate Change (IPCC) is to limit global temperature increase to 1.5 °C by 2050. Achieving this goal requires reducing global net CO2 emissions to zero emissions by 2050.
CCS plays a crucial role in that target by capturing and securely storing CO2 emissions, thereby preventing them from entering the atmosphere. This technology is essential for mitigating emissions from challenging sectors that need decarbonization, complementing renewable energy adoption, and enhancing overall carbon reduction strategies. On the other hand, CCU offers a promising approach to help meet that target by using CO2 by-products from industrial processes within the facilities of the capturing companies. This avoids the costs associated with emission tariffs and provides opportunities to generate valuable products, thereby opening new business lines in various industries.
Currently, the majority of captured CO2 is destined for CCS, with about 95% of the CO2 captured being geologically stored. Only a small fraction, less than 5%, is destined for CCU, primarily for producing synthetic fuels and chemicals. Nowadays, one of the primary uses for captured CO2 is EOR, as depicted in Figure 6a, mainly due to the economic return it provides.
CCU faces several challenges that hinder its widespread adoption. The primary obstacles include high initial capital investment, technological barriers, and the need for substantial advancements in catalyst efficiency and reactor design to achieve high conversion rates and selectivities. To make CCU economically feasible, it is essential to focus on reducing the costs of key processes, such as hydrogen production via electrolysis, by leveraging economies of scale and advancements in renewable energy technologies. Additionally, developing efficient and scalable technologies for CO2 conversion and creating supportive regulatory frameworks and financial incentives can further promote the adoption of CCU.
By 2050, the expectation is that the role of CCU will grow as technologies mature and become more economically feasible. However, CCS is still projected to be the dominant method for managing captured CO2 due to its large-scale capacity and longer-term storage solutions [5].
Figure 12 displays the technological development based on the Technology Readiness Level. The matrix shows the energy sector’s maturity of storage and utilization technologies. The TRL scale is divided into three sections: research (TRL 1–3), development (TRL 3–6), and deployment (TRL 7–9). The figure shows that Sabatier methanation and methanol production are well-understood technologies with a TRL of 9 and 8, respectively. Figure 12 shows that both CCU and CCS include technologies that have reached TRL 9. This means that these approaches are scientifically validated and ready for deployment at a commercial scale, providing reliable solutions for reducing CO2 emissions.
Some of the CCU technologies still require further research [168]. For example, the electrochemical reduction of CO2 necessitates a highly selective and efficient catalyst and a substantial energy supply. Therefore, more research is needed to better understand the catalyst’s working mechanism, degradation, and morphological changes over time to enhance its longevity and efficiency [169,170]. Photocatalytic CO2 reduction faces ongoing challenges related to photocatalyst activity, scalability, and limited efficiency. Developing more advanced photocatalysts with higher activity and stability is crucial to making this process more viable on a larger scale. In the case of the biological CO2 conversion, the technology is currently hindered by slow microorganism growth rates and the inefficiency of biological processes under industrial conditions. Enhancements in microbial engineering and process optimization are necessary to increase the rate of CO2 conversion. Furthermore, advancements in the purification and separation stages are required to make these processes economically feasible and scalable.
Beyond technical challenges, regulatory approval and public acceptance are critical factors that need to be addressed to successfully implement future Carbon Capture and Utilization (CCU) systems [28]. Public perception can be influenced by clear communication of CCU technologies’ environmental and economic benefits. Developing robust regulatory frameworks that ensure safety, efficiency, and environmental protection will be essential.
Integrating CCS and CCU into a comprehensive Carbon Capture, Utilization, and Storage (CCUS) framework can address these requirements and provide innovative solutions to reduce GHG emissions. CCUS enhances the overall efficiency and effectiveness of CO2 management and supports the transition towards a circular carbon economy by converting waste CO2 into useful products. CCUS is, therefore, an integrated approach to mitigate CO2 emissions that involves the three main stages: capturing CO2 at the emission source, utilizing the captured CO2 in various industrial applications, and securely storing the remaining CO2 in geological formations.
Considering Figure 6a,b, and assuming that 33% of the global CO2 demand will be used for EOR, the amount of CO2 available for CCUS in the energy sector is projected to be about 100 Mt/yr by 2030 and approximately 140 Mt/yr by 2050.
The SNG production will increase together with the increase in global CO2 demand. Thus, the projected SNG share will slowly replace some demand for natural gas and offer the opportunity to use CO2 that would otherwise go to EOR projects. Theoretically, natural gas demand could be supplied with the methanation of approximately 8–10 Mt/yr of captured CO2, a small fraction of the current emissions. The energy transition will probably continue to rely on natural gas due to the fluctuation of the performance of clean energy technologies. Methanol would hopefully represent about 30% of total gasoline demand by 2050. Likewise, methanol and electric cars will replace gasoline engines because many countries will ban selling internal combustion engines after 2035.
Shifting from CCS technologies like EOR to CCU technologies like SNG and methanol production involves several potential challenges and limitations. The current market price disparity is a significant limitation, with hydrogen being approximately 5–6 times more expensive per kWh than synthetic natural gas. This economic disincentive makes using green hydrogen for methane or methanol production less attractive. Additionally, the initial capital investment for SNG and methanol production facilities is substantial. These projects require advanced technologies and specialized infrastructure, such as high-efficiency electrolyzers for hydrogen production and reactors for the Sabatier process or methanol synthesis, which can be cost-prohibitive. Technological barriers include the need for continuous advancements in catalyst efficiency, reactor design, and process integration to achieve high conversion rates and selectivities, which are crucial for economic viability. Furthermore, the scalability and integration of these technologies into existing energy systems pose additional challenges that must be addressed to realize their full potential.
The International Energy Agency (IEA) explores three World Energy Outlook 2022 scenarios based on government policy assumptions to meet IPCC targets. The Stated Policies Scenario (STEPS) follows the current trajectory, the Announced Pledges Scenario (APS) assumes that all government targets will be met on time, and the Net Zero Emission scenario (NZE) aims to limit the global average temperature increase to 1.5 °C [171]. Table 4 presents the global projected natural gas and gasoline demand for 2030 and 2050 under these scenarios, along with the theoretical share of SNG and methanol in the future energy market.

5. Conclusions

The technologies for CO2 capture, utilization, and storage are crucial for achieving the net-zero emission target and limiting global temperature rise to 1.5 °C by 2050. CCUS technologies offer diverse and valuable approaches to managing CO2 emissions. In addition to CCUS, other strategies are essential for meeting climate goals. These include promoting renewable energy sources, increasing energy system efficiency, and adopting non-CO2 emitting technologies such as nuclear energy. Each approach complements the others, creating a multifaceted strategy for mitigating climate change and ensuring sustainability.
The maturity of CCS technologies, particularly EOR, EGR, aquifers, and cavern storage, is well established. These technologies have proven effective and reliable for long-term CO2 sequestration. In contrast, CCU technologies, though promising, are still emerging. CCU technologies that are technically ready (TRL 9) are heavily dependent on the availability of affordable green H2. Currently, green H2 is generally more valuable than the potential CCU products (such as CH4, CH3OH, DME), which economically limits the development and widespread adoption of CCU. Addressing this economic constraint is crucial for the advancement of CCU.
The Power-to-Gas approach through the Sabatier methanation process can store substantial energy over the long term and provide ancillary services to stabilize the future Smart Energy System. By increasing the production of SNG, the share of renewable energy in the transportation, district heating, and electricity sectors can be significantly increased, facilitating a more sustainable energy transition. However, the efficiency of electrolyzers, which are essential for hydrogen production, is susceptible to fluctuations in renewable energy input. Methanol and DME production processes offer sustainable and renewable alternatives to gasoline and diesel. Both methanol and DME are commercially viable and poised to play essential roles in the future transportation sector. Additionally, Fischer–Tropsch fuels present various possibilities for producing gasoline, diesel, and even jet fuel through renewable pathways. Dimethyl carbonate is another promising chemical with multiple applications, including its use as an electrolyte in lithium-ion batteries.
The future energy system must integrate CCS and CCU technologies to reduce GHG emissions and to effectively ensure energy supply. This integration facilitates a more comprehensive approach to managing CO2, including capturing, converting, storing, and reusing CO2. CCUS must address several technical and economic challenges. Key among these is the integration of renewable technologies to ensure a constant flow of affordable green hydrogen for utilization routes.
Further research is required to establish effective performance indicators based on techno-economic analyses and Life Cycle Assessments (LCAs) to evaluate CCUS technologies comprehensively and scientifically. These efforts are critical for overcoming existing challenges and advancing the deployment of these technologies.

Author Contributions

Conceptualization, M.V.-G. and J.M.P.-G.; methodology, J.A.G.; investigation, J.A.G.; writing—original draft preparation, J.A.G., M.V.-G., J.M.R.-M. and J.M.P.-G.; writing—review and editing, M.V.-G. and J.M.P.-G.; visualization, J.M.R.-M.; supervision, J.M.P.-G. and M.V.-G.; project administration, M.V.-G. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the grant from the TED2021-130756B-C31 funded by the MCIN/AEI/10.13039/501100011033 and by the program “ERDF A way of making Europe” by the European Union NextGenerationEU/PRTR. The authors acknowledge the project “Evaluation and Competence Development for Blue Entrepreneurship” PY20_00933.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable.

Conflicts of Interest

Author Jose Antonio Garcia was employed by the company Carbotech Gas Systems GmbH. 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.

Abbreviations

The following abbreviations are used in this manuscript:
APSAnnounced Pledges Scenario
AUDAustralian dollars
BOEMUS Bureau of Ocean Energy Management
CAESCompressed Air Energy Storage
CAPEXCapital Expenditure
CBMRCoal Bed Methane Recovery
CCSCarbon Capture and Storage
CCUCarbon Capture and Utilization
CEEPCritical Excess Electricity Production
DMCDimethyl Carbonate
DMEDimethyl Ether
ECBMREnhanced Coal bed Methane Recovery
EGREnhanced Gas Recovery
EOREnhanced Oil Recovery
ETSEuropean Emissions Trading Systems
IPCCIntergovernmental Panel on Climate Change
IEAInternational Energy Agency
LAESLiquid Air Energy Storage
LCALife Cycle Asestment
LNGLiquid Natural Gas
MCFCsMolten Carbonate Fuel cells
NZDNew Zealand Dollar
NZENet Zero Emission scenario
OPEXOperating Expediture
PHSPumped hydro storage
RESRenewable Energy Sources
RWGSReverse Water Gas Shift
SEPSurplus of Electricity Production
SNGSynthetic Natural Gas
SOECSolid Oxide Electrolysis Cell
STEPSStated Policies Scenario
TRLTechnology Readiness Level
UNEPUN Environment Program
UNFCCCUN Framework Convention on Climate Change

References

  1. Baker, H.S.; Millar, R.J.; Karoly, D.J.; Beyerle, U.; Guillod, B.P.; Mitchell, D.; Shiogama, H.; Sparrow, S.; Woollings, T. Higher CO2 concentrations increase extreme event risk in a 1.5 °C world. Nature 2018, 8, 604–608. [Google Scholar] [CrossRef]
  2. Brook, E.J.; Buizert, C. Antarctic and global climate history viewed from ice cores. Nature 2018, 558, 200–208. [Google Scholar] [CrossRef]
  3. Seneviratne, S.I.; Rogelj, J.; Séférian, R.; Wartenburger, R.; Allen, M.R.; Cain, M.; Millar, R.J.; Ebi, K.L.; Ellis, N.; Hoegh-Guldberg, O.; et al. The many possible climates from the Paris Agreement’s aim of 1.5 °C warming. Nature 2018, 558, 41–49. [Google Scholar] [CrossRef]
  4. European Commission. Paris Agreement, Climate Action; Technical Report 4; European Commission: Luxembourg, 2017. [Google Scholar]
  5. International Energy Agency. CO2 Emissions in 2023—A New Record High, but Is There Light at the End of the Tunnel? Licence: CC BY 4.0; Technical Report; International Energy Agency: Paris, France, 2024. [Google Scholar]
  6. Huang, M.T.; Zhai, P.M. Achieving Paris Agreement temperature goals requires carbon neutrality by middle century with far-reaching transitions in the whole society. Adv. Clim. Chang. Res. 2021, 12, 281–286. [Google Scholar] [CrossRef]
  7. IPCC. IPCC Special Report on Impacts of Global Warming of 1.5 °C above Pre-Industrial Levels in Context of Strengthening Response to Climate Change, Sustainable Development, and Efforts to Eradicate Poverty; Technical Report; IPCC: Geneva, Switzerland, 2022. [Google Scholar]
  8. Masood, E.; Tollefson, J. COP26 hasn’t solved the problem’: Scientists react to UN climate deal. Nature 2021, 599, 355–356. [Google Scholar] [CrossRef]
  9. Baylin-Stern, A.; Berghout, N. Is Carbon Capture too Expensive? Technical Report; International Energy Agency: Paris, France, 2021. [Google Scholar]
  10. United Nations Environment Programme. Emissions Gap Report 2022: The Closing Window—Climate Crisis Calls for Rapid Transformation of Societies; Technical Report; United Nations Environment Programme: Nairobi, Kenya, 2022. [Google Scholar]
  11. UNDP. Brief on COP27 Outcomes and Roadmap to COP28; Technical Report; United Nations Development Programme: New York, NY, USA, 2023. [Google Scholar]
  12. International Energy Agency. Supporting and Reporting Global Climate Progress. COP28: Tracking the Energy Outcomes; Technical Report; International Energy Agency: Paris, France, 2024. [Google Scholar]
  13. Feron, P.H.M.; Hendriks, C.A. CO2 Capture Process Principles and Costs. Oil Gas Sci. Technol. 2005, 60, 451–459. [Google Scholar] [CrossRef]
  14. European-Commission. The Potential for CCS and CCU in Europe; Technical Report; European-Commission: Luxembourg, 2019. [Google Scholar]
  15. Luo, J.; Xie, Y.; Hou, M.Z.; Xiong, Y.; Wu, X.; Lüddeke, C.T.; Huang, L. Advances in subsea carbon dioxide utilization and storage. Energy Rev. 2023, 2, 100016. [Google Scholar] [CrossRef]
  16. Pollitt, M.G.; von der Fehr, N.H.M.; Willems, B.; Banet, C.; Coq, C.L.; Chyong, C.K. Recommendations for a future-proof electricity market design in Europe in light of the 2021–23 energy crisis. Energy Policy 2024, 188, 114051. [Google Scholar] [CrossRef]
  17. Lund, H.; Münster, E. Management of surplus electricity-production from a fluctuating renewable-energy source. Appl. Energy 2003, 76, 65–74. [Google Scholar] [CrossRef]
  18. Koytsoumpa, E.I.; Bergins, C.; Kakaras, E. The CO2 economy: Review of CO2 capture and reuse technologies. J. Supercrit. Fluids 2018, 132, 3–16. [Google Scholar] [CrossRef]
  19. Rajabloo, T.; Valee, J.; Marenne, Y.; Coppens, L.; De Ceuninck, W. Carbon capture and utilization for industrial applications. Energy Rep. 2023, 9, 111–116. [Google Scholar] [CrossRef]
  20. Rubin, E.S.; Davison, J.E.; Herzog, H.J. The cost of CO2 capture and storage. Int. J. Greenh. Gas Control 2015, 40, 378–400. [Google Scholar] [CrossRef]
  21. Carbon-Credits. Live Carbon Prices Today; Carbon-Credits: Houston, TX, USA, 2024. [Google Scholar]
  22. Leonzio, G. State of art and perspectives about the production of methanol, dimethyl ether and syngas by carbon dioxide hydrogenation. J. CO2 Util. 2018, 27, 326–354. [Google Scholar] [CrossRef]
  23. Garcia, J.A.; Villen-Guzman, M.; Rodriguez-Maroto, J.M.; Paz-Garcia, J.M. Technical analysis of CO2 capture pathways and technologies. J. Environ. Chem. Eng. 2022, 10, 108470. [Google Scholar] [CrossRef]
  24. McLaughlin, H.; Littlefield, A.A.; Menefee, M.; Kinzer, A.; Hull, T.; Sovacool, B.K.; Bazilian, M.D.; Kim, J.; Griffiths, S. Carbon capture utilization and storage in review: Sociotechnical implications for a carbon reliant world. Renew. Sustain. Energy Rev. 2023, 177, 113215. [Google Scholar] [CrossRef]
  25. Malischek, R.; McCulloch, S. The World Has Vast Capacity to Store CO2: Net Zero Means We’ll Need It; IEA: Paris, France, 2021. [Google Scholar]
  26. Global-CCS-Institute. Global Status of CCS 2022; Technical Report; Global-CCS-Institute: Docklands, VIC, Australia, 2022. [Google Scholar]
  27. Solomon, S.; Carpenter, M.; Flach, T.A. Intermediate storage of carbon dioxide in geological formations: A technical perspective. Int. J. Greenh. Gas Control 2008, 2, 502–510. [Google Scholar] [CrossRef]
  28. Leonzio, G.; Shah, N. Recent advancements and challenges in carbon capture, utilization and storage. Curr. Opin. Green Sustain. Chem. 2024, 46, 100895. [Google Scholar] [CrossRef]
  29. 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]
  30. Valluri, S.; Claremboux, V.; Kawatra, S. Opportunities and challenges in CO2 utilization. J. Environ. Sci. 2022, 113, 322–344. [Google Scholar] [CrossRef]
  31. Kumar, N.; Verma, A.; Ahmad, T.; Sahu, R.K.; Mandal, A.; Mubashir, M.; Ali, M.; Pal, N. Carbon capture and sequestration technology for environmental remediation: A CO2 utilization approach through EOR. Geoenergy Sci. Eng. 2024, 234, 212619. [Google Scholar] [CrossRef]
  32. Chen, S.; Li, J.; Chen, S.; Qin, J.; Wang, G.; Yi, Y.; Zhang, L.; Wang, R.; Zhu, D. Experimental Evaluation of Combined Plugging System for CO2-Improved Oil Recovery and Storage. Energy Fuels 2023, 37, 4401–4412. [Google Scholar] [CrossRef]
  33. Sweatman, R.; Crookshank, S.; Edman, S. Outlook and Technologies for Offshore CO2 EOR/CCS Projects. Offshore Technol. Conf. Proc. 2011, 4, 2981–2993. [Google Scholar] [CrossRef]
  34. Bajpai, S.; Shreyash, N.; Singh, S.; Memon, A.R.; Sonker, M.; Tiwary, S.K.; Biswas, S. Opportunities, challenges and the way ahead for carbon capture, utilization and sequestration (CCUS) by the hydrocarbon industry: Towards a sustainable future. Energy Rep. 2022, 8, 111–116. [Google Scholar] [CrossRef]
  35. Zhang, C.; Wang, E.; Li, B.; Kong, X.; Xu, J.; Peng, S.; Chen, Y. Laboratory experiments of CO2-enhanced coalbed methane recovery considering CO2 sequestration in a coal seam. Energy 2023, 262, 125473. [Google Scholar] [CrossRef]
  36. Liang, W.; Wang, J.; Sang, S.; Li, P. The influence of closed pores and stacked coal grains on gas transport in CO2 injection enhanced CH4 recovery process. J. Pet. Sci. Eng. 2022, 212, 110303. [Google Scholar] [CrossRef]
  37. White, C.M.; Strazisar, B.R.; Granite, E.J.; Hoffman, J.S.; Pennline, H.W. Separation and Capture of CO2 from Large Stationary Sources and Sequestration in Geological Formations—Coalbeds and Deep Saline Aquifers. J. Air Waste Manag. Assoc. 2012, 53, 645–715. [Google Scholar] [CrossRef]
  38. Zhou, L.; Zheng, X.; Lu, Y.; Li, H.; Feng, M. Fracture pattern and caprock integrity analyses via hydraulic fracturing for CO2 enhanced coal bed methane. Eng. Fract. Mech. 2020, 228, 106894. [Google Scholar] [CrossRef]
  39. Thomsen, S.; Flørning, J. CO2 Neutral Energy System Utilising the Subsurface; Technical Report; Energy Technology Development and Demonstration Programme (EUDP): Copenhagen, Denmark, 2019. [Google Scholar]
  40. Crotogino, F.; Prelicz, R.; Rudolph, T. Assessment of the potential, the actors and relevant business cases for large scale and seasonal storage of renewable electricity by hydrogen underground storage in Europe. In Overview on All Known Underground Storage Technologies for Hydrogen Status; Technical Report; 2014; D(4). Grant Agreement No. 303417. Funded under Specific Programme “Cooperation”: Joint Technology Initiatives; Available online: https://cordis.europa.eu/project/id/303417/reporting (accessed on 26 July 2024).
  41. Chen, Y.; Chen, S.; Li, D.; Jiang, X. Density-Driven Convection for CO2 Solubility Trapping in Saline Aquifers: Modeling and Influencing Factors. Geotechnics 2023, 3, 70–103. [Google Scholar] [CrossRef]
  42. Yan, S.; Yadong, L.; Yushi, R.; Ze, B.; Hao, B.; Mingqi, L.; Qingchen, Z. Geochemical reaction of compressed CO2 energy storage using saline aquifer. Alex. Eng. J. 2023, 54, 679–689. [Google Scholar] [CrossRef]
  43. Anthonsen, K.; Aagaard, P.; Bergmo, P.; Erlström, M.; Fareide, J.; Gislason, S.; Mortensen, G.; Snæbjörnsdottir, S. CO2 storage potential in the Nordic region. Energy Procedia 2013, 37, 5080–5092. [Google Scholar] [CrossRef]
  44. Kaufmann, R.; Aavatsmark, I.; Nøkleby, P.H.; Aurdal, T.A. Using an aquifer as CO2 buffer storage. Int. J. Greenh. Gas Control 2016, 53, 106–116. [Google Scholar] [CrossRef]
  45. Mokhatab, S.; Poe, W.; Mak, J. Handbook of Natural Gas Transmission and Processing, 2nd ed.; Elsevier: Amsterdam, The Netherlands, 2012; pp. 1–35. [Google Scholar] [CrossRef]
  46. Bordenave, S.; Chatterjee, I.; Voordouw, G. Microbial community structure and microbial activities related to CO2 storage capacities of a salt cavern. Int. Biodeterior. Biodegrad. 2013, 81, 82–87. [Google Scholar] [CrossRef]
  47. Bachu, S. Sequestration of CO2 in geological media: Criteria and approach for site selection in response to climate change. Energy Convers. Manag. 2000, 41, 953–970. [Google Scholar] [CrossRef]
  48. Rackley, S.A. Carbon Capture and Storage, 2nd ed.; Butterworth-Heinemann: Oxford, UK, 2017. [Google Scholar] [CrossRef]
  49. Gunter, W.; Bachu, S.; Palombi, D.; Lakeman, B.; Sawchuk, W.; Bonner, D. Heartland Area Redwater reef saline aquifer CO2 storage project. Energy Procedia 2009, 1, 3943–3950. [Google Scholar] [CrossRef]
  50. Reichle, D.E. The Global Carbon Cycle and Climate Change—Scaling Ecological Energetics from Organism to the Biosphere; Elsevier: Amsterdam, The Netherlands, 2023. [Google Scholar] [CrossRef]
  51. Bergmo, P.S.; Grimstad, A.A.; Lindeberg, E.; Riis, F.; Johansen, W.T. Exploring geological storage sites for CO2 from Norwegian gas power plants: Utsira South. Energy Procedia 2009, 1, 2953–2959. [Google Scholar] [CrossRef]
  52. Korbøl, R.; Kaddour, A. Sleipner vest CO2 disposal—Injection of removed CO2 into the utsira formation. Energy Convers. Manag. 1995, 36, 509–512. [Google Scholar] [CrossRef]
  53. Kongsjorden, H.; Karstad, O.; Torp, T.A. Saline aquifer storage of carbon dioxide in the Sleipner project. Waste Manag. 1998, 17, 303–308. [Google Scholar] [CrossRef]
  54. Hooper, B.N. The Latrobe Valley post combustion capture project. Energy Procedia 2009, 1, 1367–1372. [Google Scholar] [CrossRef]
  55. Hjelm, L.; Anthonsen, K.L.; Dideriksen, K.; Nielsen, C.M.; Nielsen, L.H.; Mathiesen, A. Capture, Storage and Use of CO2 (CCUS)—Evaluation of the CO2 Storage Potential in Denmark; Technical Report; Geological Survey of Denmark and Greenland: København, Denmark, 2020. [Google Scholar]
  56. Mohammadian, E.; Hadavimoghaddam, F.; Kheirollahi, M.; Jafari, M.; Liu, B. Probing Solubility and pH of CO2 in aqueous solutions: Implications for CO2 injection into oceans. J. CO2 Util. 2023, 71, 102463. [Google Scholar] [CrossRef]
  57. Keil, R.G.; Nuwer, J.M.; Strand, S.E. Burial of agricultural byproducts in the deep sea as a form of carbon sequestration: A preliminary experiment. Mar. Chem. 2010, 122, 91–95. [Google Scholar] [CrossRef]
  58. Sheps, K.M.; Max, M.D.; Osegovic, J.P.; Tatro, S.R.; Brazel, L.A. A case for deep-ocean CO2 sequestration. Energy Procedia 2009, 1, 4961–4968. [Google Scholar] [CrossRef]
  59. Seibel, B.A.; Walsh, P.J. Carbon cycle: Potential impacts of CO2 injection on deep-sea biota. Science 2001, 294, 319–320. [Google Scholar] [CrossRef]
  60. Masson-Delmotte, V.; Zhai, P.; Pirani, A.; Connors, S.; Péan, C.; Berger, S.; Caud, N.; Chen, Y.; Goldfarb, L.; Gomis, M.; et al. Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, UK, 2021. [Google Scholar]
  61. Global-CCS-Institute. Global Status of CCS 2021—CCS Accelerating the Net Zero; Technical Report; Global-CCS-Institute: Docklands, VIC, Australia, 2021. [Google Scholar]
  62. 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]
  63. IEA. Putting CO2 to Use—Creating Value from Emissions; Technical Report; IEA: Paris, France, 2019. [Google Scholar]
  64. Bocin-Dumitriu, A.; Perez-Fortes, M.; Tzimas, E.; Sveen, T. Carbon Capture and Utilization Workshop—Background and Proceedings; Technical Report; European Commission: Luxembourg, 2013. [Google Scholar] [CrossRef]
  65. Global CCS Institute. Accelerating the Uptake of CCS: Industrial Use of Captured Carbon Dioxide; Technical Report; Global CCS Institute: Docklands, VIC, Australia, 2011. [Google Scholar]
  66. A Roadmap for the Global Implementation of Carbon Utilization Technologies—Transforming CO2 from a Liability to an Asset at Significant Market Scale; Technical Report; The Global CO2 Initiative (GCI) and CO2 Sciences. 2016. Available online: https://assets.ctfassets.net/xg0gv1arhdr3/5VPLtRFY3YAIasum6oYkaU/48b0f48e32d6f468d71cd80dbd451a3a/CBPI_Roadmap_Executive_Summary_Nov_2016_web.pdf (accessed on 15th June 2024).
  67. Stolaroff, J.K.; Lowry, G.V.; Keith, D.W. Using CaO- and MgO-rich industrial waste streams for carbon sequestration. Energy Convers. Manag. 2005, 46, 687–699. [Google Scholar] [CrossRef]
  68. Stolaroff, J.K.; Keith, D.W.; Lowry, G.V. Carbon Dioxide Capture from Atmospheric Air Using Sodium Hydroxide Spray. Environ. Sci. Technol. 2008, 42, 2728–2735. [Google Scholar] [CrossRef] [PubMed]
  69. Lackner, K.S. The thermodynamics of direct air capture of carbon dioxide. Energy 2013, 50, 38–46. [Google Scholar] [CrossRef]
  70. Zeman, F. Energy and Material Balance of CO2 Capture from Ambient Air. Environ. Sci. Technol. 2007, 41, 7558–7563. [Google Scholar] [CrossRef] [PubMed]
  71. Baciocchi, R.; Storti, G.; Mazzotti, M. Process design and energy requirements for the capture of carbon dioxide from air. Chem. Eng. Process. Process Intensif. 2006, 45, 1047–1058. [Google Scholar] [CrossRef]
  72. Ranjan, M.; Herzog, H.J. Feasibility of air capture. Energy Procedia 2011, 4, 2869–2876. [Google Scholar] [CrossRef]
  73. National Energy Technology Laboratory. Carbon Transport & Storage Program; Technical Report; US Department of Energy: Washington, DC, USA, 2013.
  74. Georgakaki, A.; Sales Agut, C.; Tzimas, E.; Perez Fortes, M.d.M. SETIS Magazine: Carbon Capture Utilisation and Storage; Technical Report; European Commission: Luxembourg, 2016. [Google Scholar]
  75. Schüwer, D.; Arnold, K.; Bienge, K.; Bringezu, S.; Echternacht, L.; Esken, A.; Fischedick, M.; von Geibler, J.; Höller, S.; Merten, F.; et al. CO2 Reuse NRW: Evaluating Gas Sources, Demand and Utilization for CO2 and H2 within the North Rhine-Westphalia Area with Respect to Gas Qualities; Technical Report; Wuppertal Institute for Climate, Environment and Energy: Wuppertal, Germany, 2015. [Google Scholar]
  76. Styring, P.; Jansen, D.; de Coninck, H.; Reith, H.; Armstrong, K. Carbon Capture and Utilisation in the Green Economy Using CO2 to Manufacture Fuel, Chemicals and Materials; Technical Report; The Centre for Low Carbon Futures & CO2Chem: New York, NY, USA, 2011. [Google Scholar]
  77. Vilbergsson, K.V.; Dillman, K.; Emami, N.; Ásbjörnsson, E.J.; Heinonen, J.; Finger, D.C. Can remote green hydrogen production play a key role in decarbonizing Europe in the future? A cradle-to-gate LCA of hydrogen production in Austria, Belgium, and Iceland. Int. J. Hydrogen Energy 2023, 48, 17711–17728. [Google Scholar] [CrossRef]
  78. Billig, E.; Decker, M.; Benzinger, W.; Ketelsen, F.; Pfeifer, P.; Peters, R.; Stolten, D.; Thrän, D. Non-fossil CO2 recycling—The technical potential for the present and future utilization for fuels in Germany. J. CO2 Util. 2019, 30, 130–141. [Google Scholar] [CrossRef]
  79. Nielsen, S.; Skov, I.R. Investment screening model for spatial deployment of power-to-gas plants on a national scale—A Danish case. Int. J. Hydrogen Energy 2019, 44, 9544–9557. [Google Scholar] [CrossRef]
  80. Ghaib, K.; Ben-Fares, F.Z. Power-to-Methane: A state-of-the-art review. Renew. Sustain. Energy Rev. 2018, 81, 433–446. [Google Scholar] [CrossRef]
  81. Brynolf, S.; Taljegard, M.; Grahn, M.; Hansson, J. Electrofuels for the transport sector: A review of production costs. Renew. Sustain. Energy Rev. 2018, 81, 1887–1905. [Google Scholar] [CrossRef]
  82. Jürgensen, L.; Ehimen, E.A.; Born, J.; Holm-Nielsen, J.B. Dynamic biogas upgrading based on the Sabatier process: Thermodynamic and dynamic process simulation. Bioresour. Technol. 2015, 178, 323–329. [Google Scholar] [CrossRef]
  83. Müller, K.; Städter, M.; Rachow, F.; Hoffmannbeck, D.; Schmeißer, D. Sabatier-based CO2-methanation by catalytic conversion. Environ. Earth Sci. 2013, 70, 3771–3778. [Google Scholar] [CrossRef]
  84. Stangeland, K.; Kalai, D.; Li, H.; Yu, Z. CO2 Methanation: The Effect of Catalysts and Reaction Conditions. Energy Procedia 2017, 105, 2022–2027. [Google Scholar] [CrossRef]
  85. Benjaminsson, G.; Benjaminsson, J.; Rudberg, R.B. Power-to-Gas—A Technical Review. 2013. Available online: https://www.semanticscholar.org/paper/Power-to-Gas-%E2%80%93-A-technical-review-Benjaminsson-Benjaminsson/c87e86c42959ed1879f85e8f78c715aaf78b7e23 (accessed on 15th June 2024).
  86. Leonzio, G. Process analysis of biological Sabatier reaction for bio-methane production. Chem. Eng. J. 2016, 290, 490–498. [Google Scholar] [CrossRef]
  87. Götz, M.; Lefebvre, J.; Mörs, F.; McDaniel, A.; Graf, F.; Bajohr, S.; Reimert, R.; Kolb, T. Renewable Power-to-Gas: A technological and economic review. Renew. Energy 2016, 85, 1371–1390. [Google Scholar] [CrossRef]
  88. Schlereth, D.; Hinrichsen, O. A fixed-bed reactor modeling study on the methanation of CO2. Chem. Eng. Res. Design 2014, 92, 702–712. [Google Scholar] [CrossRef]
  89. Rönsch, S.; Schneider, J.; Matthischke, S.; Schlüter, M.; Götz, M.; Lefebvre, J.; Prabhakaran, P.; Bajohr, S. Review on methanation—From fundamentals to current projects. Fuel 2016, 166, 276–296. [Google Scholar] [CrossRef]
  90. Bailera, M.; Lisbona, P.; Peña, B.; Alarcón, A.; Guilera, J.; Perpiñán, J.; Romeo, L.M. Synthetic natural gas production in a 1 kW reactor using Ni–Ce/Al2O3 and Ru–Ce/Al2O3: Kinetics, catalyst degradation and process design. Energy 2022, 256, 124720. [Google Scholar] [CrossRef]
  91. Guilera, J.; Ramon Morante, J.; Andreu, T. Economic viability of SNG production from power and CO2. Energy Convers. Manag. 2018, 162, 218–224. [Google Scholar] [CrossRef]
  92. Sveinbjörnsson, D.; Münster, E.; Aryal, N.; Bo, R.; Pedersen, B. WP1 Gas Conditioning and Grid Operation—Upgrading of Biogas to Biomethane with the Addition of Hydrogen from Electrolysis; Technical Report; PlanEnergi: København, Denmark, 2017. [Google Scholar]
  93. Corbellini, V.; Kougias, P.G.; Treu, L.; Bassani, I.; Malpei, F.; Angelidaki, I. Hybrid biogas upgrading in a two-stage thermophilic reactor. Energy Convers. Manag. 2018, 168, 1–10. [Google Scholar] [CrossRef]
  94. Luo, G.; Angelidaki, I. Integrated biogas upgrading and hydrogen utilization in an anaerobic reactor containing enriched hydrogenotrophic methanogenic culture. Biotechnol. Bioeng. 2012, 109, 2729–2736. [Google Scholar] [CrossRef] [PubMed]
  95. Luo, G.; Johansson, S.; Boe, K.; Xie, L.; Zhou, Q.; Angelidaki, I. Simultaneous hydrogen utilization and in situ biogas upgrading in an anaerobic reactor. Biotechnol. Bioeng. 2012, 109, 1088–1094. [Google Scholar] [CrossRef] [PubMed]
  96. Juan-Juan, J.; Román-Martínez, M.C.; Illán-Gómez, M.J. Nickel catalyst activation in the carbon dioxide reforming of methane: Effect of pretreatments. Appl. Catal. A Gen. 2009, 355, 27–32. [Google Scholar] [CrossRef]
  97. Kim, J.H.; Suh, D.J.; Park, T.J.; Kim, K.L. Effect of metal particle size on coking during CO2 reforming of CH4 over Ni–alumina aerogel catalysts. Appl. Catal. A Gen. 2000, 197, 191–200. [Google Scholar] [CrossRef]
  98. Nikoo, M.K.; Amin, N.A. Thermodynamic analysis of carbon dioxide reforming of methane in view of solid carbon formation. Fuel Process. Technol. 2011, 92, 678–691. [Google Scholar] [CrossRef]
  99. Damyanova, S.; Pawelec, B.; Arishtirova, K.; Fierro, J.L. Ni-based catalysts for reforming of methane with CO2. Int. J. Hydrogen Energy 2012, 37, 15966–15975. [Google Scholar] [CrossRef]
  100. Guharoy, U.; Le Saché, E.; Cai, Q.; Reina, T.R.; Gu, S. Understanding the role of Ni-Sn interaction to design highly effective CO2 conversion catalysts for dry reforming of methane. J. CO2 Util. 2018, 27, 1–10. [Google Scholar] [CrossRef]
  101. Kaydouh, M.N.; El Hassan, N.; Davidson, A.; Casale, S.; El Zakhem, H.; Massiani, P. Effect of the order of Ni and Ce addition in SBA-15 on the activity in dry reforming of methane. C. R. Chim. 2015, 18, 293–301. [Google Scholar] [CrossRef]
  102. Sokolov, S.; Kondratenko, E.V.; Pohl, M.M.; Barkschat, A.; Rodemerck, U. Stable low-temperature dry reforming of methane over mesoporous La2O3-ZrO2 supported Ni catalyst. Appl. Catal. B Environ. 2012, 113–114, 19–30. [Google Scholar] [CrossRef]
  103. Rezaei, M.; Alavi, S.M.; Sahebdelfar, S.; Yan, Z.F. Syngas Production by Methane Reforming with Carbon Dioxide on Noble Metal Catalysts. J. Nat. Gas Chem. 2006, 15, 327–334. [Google Scholar] [CrossRef]
  104. Nair, M.M.; Abanades, S. Tailoring Hybrid Nonstoichiometric Ceria Redox Cycle for Combined Solar Methane Reforming and Thermochemical Conversion of H2O/CO2. Energy Fuels 2016, 30, 6050–6058. [Google Scholar] [CrossRef]
  105. Chueh, W.C.; Haile, S.M. A thermochemical study of ceria: Exploiting an old material for new modes of energy conversion and CO2 mitigation. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2010, 368, 3269–3294. [Google Scholar] [CrossRef]
  106. Bhosale, R.R.; Takalkar, G.D. Nanostructured co-precipitated Ce0.9Ln0.1O2 (Ln = La, Pr, Sm, Nd, Gd, Tb, Dy, or Er) for thermochemical conversion of CO2. Ceram. Int. 2018, 44, 16688–16697. [Google Scholar] [CrossRef]
  107. Da Silva, R.J.; Pimentel, A.F.; Monteiro, R.S.; Mota, C.J. Synthesis of methanol and dimethyl ether from the CO2 hydrogenation over Cu·ZnO supported on Al2O3 and Nb2O5. J. CO2 Util. 2016, 15, 83–88. [Google Scholar] [CrossRef]
  108. Schemme, S.; Breuer, J.L.; Samsun, R.C.; Peters, R.; Stolten, D. Promising catalytic synthesis pathways towards higher alcohols as suitable transport fuels based on H2 and CO2. J. CO2 Util. 2018, 27, 223–237. [Google Scholar] [CrossRef]
  109. Jadhav, S.G.; Vaidya, P.D.; Bhanage, B.M.; Joshi, J.B. Catalytic carbon dioxide hydrogenation to methanol: A review of recent studies. Chem. Eng. Res. Des. 2014, 92, 2557–2567. [Google Scholar] [CrossRef]
  110. Danish Energy Agency. Technology Data for Renewable Fuels; Technical Report; Danish Energy Agency: København, Denmark, 2017. [Google Scholar]
  111. Bargiacchi, E.; Antonelli, M.; Desideri, U. A comparative assessment of Power-to-Fuel production pathways. Energy 2019, 183, 1253–1265. [Google Scholar] [CrossRef]
  112. Schemme, S.; Breuer, J.L.; Köller, M.; Meschede, S.; Walman, F.; Samsun, R.C.; Peters, R.; Stolten, D. H2-based synthetic fuels: A techno-economic comparison of alcohol, ether and hydrocarbon production. Int. J. Hydrogen Energy 2020, 45, 5395–5414. [Google Scholar] [CrossRef]
  113. Ridjan, I.; Mathiesen, B.V.; Connolly, D.; Duić, N. The feasibility of synthetic fuels in renewable energy systems. Energy 2013, 57, 76–84. [Google Scholar] [CrossRef]
  114. Decker, M.; Schorn, F.; Samsun, R.C.; Peters, R.; Stolten, D. Off-grid power-to-fuel systems for a market launch scenario—A techno-economic assessment. Appl. Energy 2019, 250, 1099–1109. [Google Scholar] [CrossRef]
  115. Bozzano, G.; Manenti, F. Efficient methanol synthesis: Perspectives, technologies and optimization strategies. Prog. Energy Combust. Sci. 2016, 56, 71–105. [Google Scholar] [CrossRef]
  116. Schmidt, P.; Weindorf, W.; Roth, A.; Batteiger, V.; Riegel, F. Power-to-Liquids Potentials and Perspectives for the Future Supply of Renewable Aviation Fuel; Technical Report; German Environment Agency: Dessau-Roßlau, Germany, 2016. [Google Scholar]
  117. CRI. Projects: Emissions-to-Liquids Technology; CRI: Tokyo, Japan, 2021. [Google Scholar]
  118. Becker, W.L.; Braun, R.J.; Penev, M.; Melaina, M. Production of Fischer–Tropsch liquid fuels from high temperature solid oxide co-electrolysis units. Energy 2012, 47, 99–115. [Google Scholar] [CrossRef]
  119. Schemme, S.; Samsun, R.C.; Peters, R.; Stolten, D. Power-to-fuel as a key to sustainable transport systems—An analysis of diesel fuels produced from CO2 and renewable electricity. Fuel 2017, 205, 198–221. [Google Scholar] [CrossRef]
  120. Ail, S.S.; Dasappa, S. Biomass to liquid transportation fuel via Fischer Tropsch synthesis —Technology review and current scenario. Renew. Sustain. Energy Rev. 2016, 58, 267–286. [Google Scholar] [CrossRef]
  121. Todic, B.; Nowicki, L.; Nikacevic, N.; Bukur, D.B. Fischer–Tropsch synthesis product selectivity over an industrial iron-based catalyst: Effect of process conditions. Catal. Today 2016, 261, 28–39. [Google Scholar] [CrossRef]
  122. Zhao, K.; Bkour, Q.; Hou, X.; Kang, S.W.; Park, J.C.; Norton, M.G.; Yang, J.I.; Ha, S. Reverse water gas shift reaction over CuFe/Al2O3 catalyst in solid oxide electrolysis cell. Chem. Eng. J. 2018, 336, 20–27. [Google Scholar] [CrossRef]
  123. Shafer, W.D.; Gnanamani, M.K.; Graham, U.M.; Yang, J.; Masuku, C.M.; Jacobs, G.; Davis, B.H. Fischer—Tropsch: Product Selectivity—The Fingerprint of Synthetic Fuels. Catalysts 2019, 9, 259. [Google Scholar] [CrossRef]
  124. International Energy Agency-AMF. Diesel and Gasoline; International Energy Agency-AMF: Paris, France, 2021. [Google Scholar]
  125. Gill, S.S.; Tsolakis, A.; Dearn, K.D.; Rodríguez-Fernández, J. Combustion characteristics and emissions of Fischer–Tropsch diesel fuels in IC engines. Prog. Energy Combust. Sci. 2011, 37, 503–523. [Google Scholar] [CrossRef]
  126. Dry, M.E. High quality diesel via the Fischer–Tropsch process—A review. J. Chem. Technol. Biotechnol. 2002, 77, 43–50. [Google Scholar] [CrossRef]
  127. Kim, Y.D.; Yang, C.W.; Kim, B.J.; Moon, J.H.; Jeong, J.Y.; Jeong, S.H.; Lee, S.H.; Kim, J.H.; Seo, M.W.; Lee, S.B.; et al. Fischer–tropsch diesel production and evaluation as alternative automotive fuel in pilot-scale integrated biomass-to-liquid process. Appl. Energy 2016, 180, 301–312. [Google Scholar] [CrossRef]
  128. Rane, S.; Borg, Ø.; Yang, J.; Rytter, E.; Holmen, A. Effect of alumina phases on hydrocarbon selectivity in Fischer–Tropsch synthesis. Appl. Catal. A Gen. 2010, 388, 160–167. [Google Scholar] [CrossRef]
  129. Sage, V.; Sun, Y.; Hazewinkel, P.; Bhatelia, T.; Braconnier, L.; Tang, L.; Chiang, K.; Batten, M.; Burke, N. Modified product selectivity in Fischer-Tropsch synthesis by catalyst pre-treatment. Fuel Process. Technol. 2017, 167, 183–192. [Google Scholar] [CrossRef]
  130. Mißbach, H.; Schmidt, B.C.; Duda, J.P.; Lünsdorf, N.K.; Goetz, W.; Thiel, V. Assessing the diversity of lipids formed via Fischer-Tropsch-type reactions. Org. Geochem. 2018, 119, 110–121. [Google Scholar] [CrossRef]
  131. Schulz, H. Short history and present trends of Fischer–Tropsch synthesis. Appl. Catal. A Gen. 1999, 186, 3–12. [Google Scholar] [CrossRef]
  132. Stranges, A. Germany’s synthetic fuel industry, 1927–1945. In The German Chemical Industry in the Twentieth Century; Springer: Dordrecht, The Netherlands, 2000; pp. 147–216. [Google Scholar] [CrossRef]
  133. Jarvis, S.M.; Samsatli, S. Technologies and infrastructures underpinning future CO2 value chains: A comprehensive review and comparative analysis. Renew. Sustain. Energy Rev. 2018, 85, 46–68. [Google Scholar] [CrossRef]
  134. Zhang, X.; Lu, H.; Miao, Y.; Zhang, Y.; Wang, J. New trends in the development of CO2 electrochemical reduction electrolyzer. J. Environ. Chem. Eng. 2024, 12, 112369. [Google Scholar] [CrossRef]
  135. Nabil, S.K.; McCoy, S.; Kibria, M.G. Comparative life cycle assessment of electrochemical upgrading of CO2 to fuels and feedstocks. Green Chem. 2021, 23, 867–880. [Google Scholar] [CrossRef]
  136. Xu, Z.; Peng, C.; Zheng, G. Coupling Value-Added Anodic Reactions with Electrocatalytic CO2 Reduction. Chem.-A Eur. J. 2023, 29, e202203147. [Google Scholar] [CrossRef] [PubMed]
  137. Fan, L.; Xia, C.; Zhu, P.; Lu, Y.; Wang, H. Electrochemical CO2 reduction to high-concentration pure formic acid solutions in an all-solid-state reactor. Nat. Commun. 2020, 11, 3633. [Google Scholar] [CrossRef] [PubMed]
  138. Roh, K.; Bardow, A.; Bongartz, D.; Burre, J.; Chung, W.; Deutz, S.; Han, D.; Heßelmann, M.; Kohlhaas, Y.; König, A.; et al. Early-stage evaluation of emerging CO2 utilization technologies at low technology readiness levels. Green Chem. 2020, 22, 3842–3859. [Google Scholar] [CrossRef]
  139. Arora, I.; Garg, S.; Sapi, A.; Ingole, P.P.; Chandra, A. Insights into photocatalytic CO2 reduction reaction pathway: Catalytic modification for enhanced solar fuel production. J. Ind. Eng. Chem. 2024, 137, 1–28. [Google Scholar] [CrossRef]
  140. Fu, J.; Jiang, K.; Qiu, X.; Yu, J.; Liu, M. Product selectivity of photocatalytic CO2 reduction reactions. Mater. Today 2020, 32, 222–243. [Google Scholar] [CrossRef]
  141. Mao, J.; Li, K.; Peng, T. Recent advances in the photocatalytic CO2 reduction over semiconductors. Catal. Sci. Technol. 2013, 3, 2481–2498. [Google Scholar] [CrossRef]
  142. Call, A.; Cibian, M.; Yamamoto, K.; Nakazono, T.; Yamauchi, K.; Sakai, K. Highly Efficient and Selective Photocatalytic CO2 Reduction to CO in Water by a Cobalt Porphyrin Molecular Catalyst. ACS Catal. 2019, 9, 4867–4874. [Google Scholar] [CrossRef]
  143. Habisreutinger, S.N.; Schmidt-Mende, L.; Stolarczyk, J.K. Photocatalytic Reduction of CO2 on TiO Semiconductors. Angew. Chem. Int. Ed. 2013, 52, 7372–7408. [Google Scholar] [CrossRef]
  144. Hasan, M.F.; First, E.L.; Boukouvala, F.; Floudas, C.A. A multi-scale framework for CO2 capture, utilization, and sequestration: CCUS and CCU. Comput. Chem. Eng. 2015, 81, 2–21. [Google Scholar] [CrossRef]
  145. Karamian, E.; Sharifnia, S. On the general mechanism of photocatalytic reduction of CO2. J. CO2 Util. 2016, 16, 194–203. [Google Scholar] [CrossRef]
  146. Vu, N.N.; Kaliaguine, S.; Do, T.O. Critical Aspects and Recent Advances in Structural Engineering of Photocatalysts for Sunlight-Driven Photocatalytic Reduction of CO2 into Fuels. Adv. Funct. Mater. 2019, 29, 1901825. [Google Scholar] [CrossRef]
  147. Abdelkareem, M.A.; Lootah, M.A.; Sayed, E.T.; Wilberforce, T.; Alawadhi, H.; Yousef, B.A.; Olabi, A.G. Fuel cells for carbon capture applications. Sci. Total Environ. 2021, 769, 144243. [Google Scholar] [CrossRef] [PubMed]
  148. Wu, J.; Chen, L. Energy Efficiency Management of Coupling System for Molten Carbonate Fuel Cells. Int. J. Electrochem. Sci. 2022, 17, 220945. [Google Scholar] [CrossRef]
  149. Suzanne, F.; Anthony, T. Molten Carbonate Fuel Cells for 90% Post Combustion CO2 Capture from a New Build CCGT. Front. Energy Res. 2021, 9, 359. [Google Scholar] [CrossRef]
  150. Azizi, Z.; Rezaeimanesh, M.; Tohidian, T.; Rahimpour, M.R. Dimethyl ether: A review of technologies and production challenges. Chem. Eng. Process. Process Intensif. 2014, 82, 150–172. [Google Scholar] [CrossRef]
  151. Michailos, S.; McCord, S.; Sick, V.; Stokes, G.; Styring, P. Dimethyl ether synthesis via captured CO2 hydrogenation within the power to liquids concept: A techno-economic assessment. Energy Convers. Manag. 2019, 184, 262–276. [Google Scholar] [CrossRef]
  152. Vad Mathiesen, B.; Johannsen, R.M.; Kermeli, K.; Crijns-Graus, W.; Lund, H.; Skov, I.R. The green transition of industry—An introduction to IndustryPLAN. Smart Energy 2023, 11, 100111. [Google Scholar] [CrossRef]
  153. Pontzen, F.; Liebner, W.; Gronemann, V.; Rothaemel, M.; Ahlers, B. CO2-based methanol and DME—Efficient technologies for industrial scale production. Catal. Today 2011, 171, 242–250. [Google Scholar] [CrossRef]
  154. Zhou, Y.J.; Xiao, M.; Wang, S.J.; Han, D.M.; Lu, Y.X.; Meng, Y.Z. Effects of Mo promoters on the Cu–Fe bimetal catalysts for the DMC formation from CO2 and methanol. Chin. Chem. Lett. 2013, 24, 307–310. [Google Scholar] [CrossRef]
  155. Zhang, M.; Alferov, K.A.; Xiao, M.; Han, D.; Wang, S.; Meng, Y. Continuous Dimethyl Carbonate Synthesis from CO2 and Methanol Using Cu-Ni@VSiO as Catalyst Synthesized by a Novel Sulfuration Method. Catalysts 2018, 8, 142. [Google Scholar] [CrossRef]
  156. Zhou, Y.; Wang, S.; Xiao, M.; Han, D.; Lu, Y.; Meng, Y. Novel Cu-Fe bimetal catalyst for the formation of dimethyl carbonate from carbon dioxide and methanol. RSC Adv. 2012, 2, 6831–6837. [Google Scholar] [CrossRef]
  157. Zhou, Y.; Fu, Z.; Wang, S.; Xiao, M.; Han, D.; Meng, Y. Electrochemical synthesis of dimethyl carbonate from CO2 and methanol over carbonaceous material supported DBU in a capacitor-like cell reactor. RSC Adv. 2016, 6, 40010–40016. [Google Scholar] [CrossRef]
  158. Selva, M.; Marques, C.A.; Tundo, P. The addition reaction of dialkyl carbonates to ketones. Gazz. Chim. Ital. 1994, 123, 515–518. [Google Scholar] [CrossRef]
  159. Bortnick, N.; Luskin, L.S.; Hurwitz, M.D.; Rytina, A.W. t-Carbinamines, RR′R″CNH2. III. The Preparation of Isocyanates, Isothiocyanates and Related Compounds1. J. Am. Chem. Soc. 2002, 78, 4358–4361. [Google Scholar] [CrossRef]
  160. Vauthey, I.; Valot, F.; Gozzi, C.; Fache, F.; Lemaire, M. An environmentally benign access to carbamates and ureas. Tetrahedron Lett. 2000, 41, 6347–6350. [Google Scholar] [CrossRef]
  161. Selva, M.; Tundo, P.; Perosa, A. The synthesis of alkyl carbamates from primary aliphatic amines and dialkyl carbonates in supercritical carbon dioxide. Tetrahedron Lett. 2002, 43, 1217–1219. [Google Scholar] [CrossRef]
  162. Cao, Y.; Cheng, H.; Ma, L.; Liu, F.; Liu, Z. Research Progress in the Direct Synthesis of Dimethyl Carbonate from CO2 and Methanol. Catal. Surv. Asia 2012, 16, 138–147. [Google Scholar] [CrossRef]
  163. Tamboli, A.H.; Chaugule, A.A.; Kim, H. Highly selective and multifunctional chitosan/ionic liquids catalyst for conversion of CO2 and methanol to dimethyl carbonates at mild reaction conditions. Fuel 2017, 166, 495–501. [Google Scholar] [CrossRef]
  164. Hong, H.N.; Samsun, R.C.; Peters, R.; Stolten, D. Greener production of dimethyl carbonate by the Power-to-Fuel concept: A comparative techno-economic analysis. Electrochem. Process. Eng. 2021, 23, 1734–1747. [Google Scholar] [CrossRef]
  165. Zhang, M.; Xu, Y.; Williams, B.L.; Xiao, M.; Wang, S.; Han, D.; Sun, L.; Meng, Y. Catalytic materials for direct synthesis of dimethyl carbonate (DMC) from CO2. J. Clean. Prod. 2021, 279, 123344. [Google Scholar] [CrossRef]
  166. Kearns, D.; Liu, H.; Consoli, C. Technology Readiness and Costs of CCS; Global CCS Institure: Docklands, VIC, Australia, 2021. [Google Scholar]
  167. Hong, W.Y. A techno-economic review on carbon capture, utilisation and storage systems for achieving a net-zero CO2 emissions future. Carbon Capture Sci. Technol. 2022, 3, 100044. [Google Scholar] [CrossRef]
  168. Bhavsar, A.; Hingar, D.; Ostwal, S.; Thakkar, I.; Jadeja, S.; Shah, M. The current scope and stand of carbon capture storage and utilization—A comprehensive review. Case Stud. Chem. Environ. Eng. 2023, 8, 100368. [Google Scholar] [CrossRef]
  169. Ahmad, T.; Liu, S.; Sajid, M.; Li, K.; Ali, M.; Liu, L.; Chen, W. Electrochemical CO2 reduction to C2+ products using Cu-based electrocatalysts: A review. Nano Res. Energy 2022, 1, 9120021. [Google Scholar] [CrossRef]
  170. Gao, Z.; Li, J.; Zhang, Z.; Hu, W. Recent advances in carbon-based materials for electrochemical CO2 reduction reaction. Chin. Chem. Lett. 2022, 33, 2270–2280. [Google Scholar] [CrossRef]
  171. IEA. World Energy Outlook 2022; Technical Report; IEA: Paris, France, 2022. [Google Scholar]
Figure 1. Number of new CCS facilities with the total CO2 storage capacity. Based on [26].
Figure 1. Number of new CCS facilities with the total CO2 storage capacity. Based on [26].
Sustainability 16 06639 g001
Figure 4. Illustration of various forms of direct CO2 injection into the oceans [56].
Figure 4. Illustration of various forms of direct CO2 injection into the oceans [56].
Sustainability 16 06639 g004
Figure 5. Global storage resources by country in late 2021, totaling a global capacity of 13,000 Gt CO2 [61,62].
Figure 5. Global storage resources by country in late 2021, totaling a global capacity of 13,000 Gt CO2 [61,62].
Sustainability 16 06639 g005
Figure 6. (a) Breakdown of global CO2 consumption in 2015. Adapted from [63]. (b) Future projection of global CO2 demand.
Figure 6. (a) Breakdown of global CO2 consumption in 2015. Adapted from [63]. (b) Future projection of global CO2 demand.
Sustainability 16 06639 g006
Figure 7. Power-to-X routes.
Figure 7. Power-to-X routes.
Sustainability 16 06639 g007
Figure 8. Methane production by Sabatier process. Adopted from [81].
Figure 8. Methane production by Sabatier process. Adopted from [81].
Sustainability 16 06639 g008
Figure 9. Direct CO2 hydrogenation for methanol production. Adopted from [81,115].
Figure 9. Direct CO2 hydrogenation for methanol production. Adopted from [81,115].
Sustainability 16 06639 g009
Figure 10. Low Fischer–Tropsch process. Adopted from [129].
Figure 10. Low Fischer–Tropsch process. Adopted from [129].
Sustainability 16 06639 g010
Figure 11. Molten carbonate fuel cell illustration with anode effluent. Based on [148].
Figure 11. Molten carbonate fuel cell illustration with anode effluent. Based on [148].
Sustainability 16 06639 g011
Figure 12. CO2 storage vs. CO2 utilization TRL matrix [29,166,167].
Figure 12. CO2 storage vs. CO2 utilization TRL matrix [29,166,167].
Sustainability 16 06639 g012
Table 1. CO2 emission tariff auction price in different markets (July 2024) [21].
Table 1. CO2 emission tariff auction price in different markets (July 2024) [21].
MarketsCarbon Price
EU ETSEUR 72.4/t
New Zealand (NZD)NZD 51.5/t
Australia (AUD)AUD 33.5/t
California (USA)USD 44.0/t
South KoreaUSD 6.5/t
ChinaUSD 12.6/t
Table 2. CO2 trapping mechanisms in saline caves [48,49,50].
Table 2. CO2 trapping mechanisms in saline caves [48,49,50].
MechanismsCO2 Trapping PhaseDescription
MineralReacted solid phaseDissolved CO2 reacts with minerals based on Fe, Ca or Mg to form carbonates
HydrodynamicSupercritical fluidUndissolved CO2 is trapped by cap rocks with low permeability
SolubilityDissolved liquid phaseCO2 is dissolved in the brine water
ResidualGas phaseCO2 displaces water from the rock pores
Table 4. Common fuels’ share in the future energy sector. “bcm/yr” stands for billion cubic meters per year and “Mb/d” stands for million barrels per day.
Table 4. Common fuels’ share in the future energy sector. “bcm/yr” stands for billion cubic meters per year and “Mb/d” stands for million barrels per day.
Year20302050
ScenariosSTEPSAPSNZESTEPSAPSNZE
Gas demand (bcm)445640693666466135682681
SNG share (%)1.121.221.361.501.962.61
Gasoline demand (mb/d)23.220.6-19.38.2-
Methanol share (%)6.67.4-11.126.2-
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

Garcia, J.A.; Villen-Guzman, M.; Rodriguez-Maroto, J.M.; Paz-Garcia, J.M. Comparing CO2 Storage and Utilization: Enhancing Sustainability through Renewable Energy Integration. Sustainability 2024, 16, 6639. https://doi.org/10.3390/su16156639

AMA Style

Garcia JA, Villen-Guzman M, Rodriguez-Maroto JM, Paz-Garcia JM. Comparing CO2 Storage and Utilization: Enhancing Sustainability through Renewable Energy Integration. Sustainability. 2024; 16(15):6639. https://doi.org/10.3390/su16156639

Chicago/Turabian Style

Garcia, Jose Antonio, Maria Villen-Guzman, Jose Miguel Rodriguez-Maroto, and Juan Manuel Paz-Garcia. 2024. "Comparing CO2 Storage and Utilization: Enhancing Sustainability through Renewable Energy Integration" Sustainability 16, no. 15: 6639. https://doi.org/10.3390/su16156639

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