The International Journal of Coal Science & Technology is a peer-reviewed open access journal. It focuses on key topics of coal scientific research and mining development, serving as a forum for scientists to present research findings and discuss challenging issues.
Coverage includes original research articles, new developments, case studies and critical reviews in all aspects of scientific and engineering research on coal, coal utilizations and coal mining. Among the broad topics receiving attention are coal geology, geochemistry, geophysics, mineralogy, and petrology; coal mining theory, technology and engineering; coal processing, utilization and conversion; coal mining environment and reclamation and related aspects.
The International Journal of Coal Science & Technology is published with China Coal Society, who also cover the publication costs so authors do not need to pay an article-processing charge.
The journal operates a single-blind peer-review system, where the reviewers are aware of the names and affiliations of the authors, but the reviewer reports provided to authors are anonymous.
A forum for new research findings, case studies and discussion of important challenges in coal science and mining development
Offers an international perspective on coal geology, coal mining, technology and engineering, coal processing, utilization and conversion, coal mining environment and reclamation and more
Published with the China Coal Society
Research Article
Open Access
Published: 18 February 2025
0 Accesses
International Journal of Coal Science & Technology Volume 12, article number 15, (2025)
1.
National Key Laboratory of Petroleum Resources and Engineering, Beijing, China
2.
Unconventional Petroleum Research Institute, China University of Petroleum, Beijing, Beijing, China
3.
National Key Laboratory of High-Efficiency Flexible Coal Power Generation and Carbon Capture Utilization and Storage, Beijing, China
4.
China Huaneng Group Clean Energy Technology Research Institute Co., Ltd., Beijing, China
The substantial emissions of greenhouse gases, particularly CO2, constitute a primary driver of global warming. CCUS is proposed as an effective mitigation strategy which is often estimated to account for about 15% of cumulative carbon emission reduction. In-situ CO2 mineralization sequestration, compared to conventional geological storage methods such as depleted oil and gas reservoirs, unmineable coal seams, and deep saline aquifers, offers the advantage of permanent immobilization of injected carbon. However, uncertainties persist regarding the characteristics of geochemical interactions under reservoir pore conditions, as well as the kinetic mechanisms of mineralization reactions. Additionally, geochemical reactions may lead to solid particle transport and deposition, potentially causing pore throat occlusion. Pilot projects in Iceland and the United States have demonstrated the feasibility of this technology, but the field remains in the early deployment stage. In this review, the mechanisms of in-situ mineralization have been elucidated, the primary factors influencing the reaction kinetics have been discussed, and the current research status in this field has been summarized. It is emphasized that establishing a reliable system for evaluating storage capacity and understanding the kinetic mechanisms governing CO2 conversion into minerals at multi-phase interfaces are key priorities for future work.
Carbon dioxide (CO2) is one of the primary greenhouse gases responsible for global climate change. As of 2022, global CO2 emissions related to the energy sector have increased to 37.4 Gt (IEA 2024). To achieve the goal of limiting global warming to no more than 1.5 °C by the end of this century, as stipulated by the Paris Agreement, the greenhouse gas emissions must be reduced by 45% compared to current levels by 2030, as reported in the Emissions Gap Report 2022 by the United Nations Environment (UNEP 2022). It is a significant challenge, hence seeking effective CO2 utilization and removal technologies is of great importance for mitigating climate change (Hepburn et al. 2019; Zhang et al. 2020). Carbon capture, utilization, and storage (CCUS) technology is widely recognized as a key approach to achieve net-zero CO2 emissions, and a viable technological option for low-carbon transformation in industries that are difficult to decarbonize, such as steel and cement (Wei et al. 2021). According to International Energy Agency (IEA), in a net-zero emissions scenario, it is anticipated that a global annual CO2 capture capacity through the implementation of CCUS technologies will reach 1.6 Gt by 2030 and increase to 7.6 Gt by 2050, making an important contribution to global carbon reduction efforts (IEA 2021). As of July 2023, there are 392 CCS facilities in various stages of development, collectively capable of capturing over 360 million metric tons of CO2 annually. Although only 49 projects are currently operational, the success of demonstration projects and improved risk monitoring have significantly increased public acceptance of CCUS technology (GCCSI 2023).
Geological carbon sequestration (GCS) is commonly implemented by injecting CO2 into geological bodies or structures within sedimentary basins, such as depleted oil and gas reservoirs, unmineable coal seams, and deep saline aquifers, to achieve permanent carbon storage (Ali et al. 2022; Aminu et al. 2017; Orr Jr 2009). There are four distinct storage mechanisms for GCS, namely structural trapping, residual trapping, dissolution trapping, and mineral trapping. Each with varying temporal dynamics and degrees of contribution and safety in CO2 storage (Metz et al. 2005). Among these, mineral trapping is considered the most secure storage mechanism because it leads to the formation of stable carbonate minerals. However, sedimentary formations are composed of complex mineral constituents, such as sandstones typically containing minerals like quartz, feldspars (potassium feldspar and sodium feldspar), carbonates (calcite, siderite, and dolomite), and clays (kaolinite, illite, smectite, and chlorite), which exhibit slow dissolution and reaction kinetics with CO2-saturated brine. Additionally, they lack the divalent metal cations necessary for forming stable carbonate minerals. These characteristics pose challenges to mineral trapping and hinder the establishment of a permanent and effective sequestration process (Jayasekara et al. 2020; De Silva and Ranjith 2012; Benson and Cole 2008).
In this scenario, the mineral carbonation of CO2 with magnesium-rich and calcium-rich silicates offers an effective alternative for GCS. CO2 mineralization simulates the natural weathering process of calcium-magnesium silicates. Seifritz 1990 first proposed that alkaline-ultramafic rock carbonate could sequester CO2, this technique has been recognized as a complementary approach for CO2 storage. It can be implemented underground (in-situ) and above ground (ex-situ) (Sanna et al. 2014). Ex-situ mineral carbonation involves reactions between captured CO2 with alkaline materials such as steel slag (Wang et al. 2023), cement (Li et al. 2022), ground reactive rocks (Sanna and Maroto-valer 2016), and mine tailings (Oskierski et al. 2021), to achieve carbonation and produce value-added industrial products (Yadav and Mehra 2021; Liu et al. 2021; Ma et al. 2022b). However, the cost of ex-situ mineral carbonation is much higher than that of CO2 geological storage, which hinders its large-scale development (Kelemen et al. 2019). On the other hand, in-situ mineral carbonation involves injecting CO2 into reactive rocks such as basalt, resulting in rapid mineral carbonation/trapping and transforming CO2 into stable carbonate minerals, greatly reducing the risk of CO2 leakage and providing a safer, unmonitored option for permanent CO2 sequestration (Sanna et al. 2014; Snæbjörnsdóttir et al. 2020; Ostovari et al. 2022). Whether it is in-situ or ex-situ mineralization, it is crucial to clarify the mineral dissolution and precipitation reactions along with their respective reaction rates, as these factors directly affect the safety, efficiency, and cost of CO2 storage.
Indeed, numerous reviews on CO2 mineralization have been published and the key details have been summarized in Table 1 (Snæbjörnsdóttir et al. 2020; Power et al. 2013; Kelemen et al. 2011; Sanna et al. 2014; Liu et al. 2021; Kim et al. 2023). However, there is a relative scarcity of reviews specifically focusing on in-situ CO2 mineralization, with greater attention seemingly focusing on the enhancement of efficiency and the optimization of process conditions of ex-situ mineralization. Power et al. (2013) summarized the fundamental principles of mineral precipitation and dissolution, highlighting that the precipitation-dissolution process was primarily governed by supersaturation and can be influenced by factors such as temperature, pH, microbial activity, or solution composition (Geerlings and Zevenhoven 2013; Lin et al. 2022, 2024a). They also emphasized that optimizing reaction conditions for direct or indirect mineral carbonation can significantly enhance reaction efficiency. Snæbjörnsdóttir et al. (2020) extensively discussed the types and distribution of rocks suitable for in-situ mineralization techniques, proposing a set of screening criteria that consider factors such as injectivity and safety (Kelemen et al. 2011; Raza et al. 2022). This has guided site selection in in-situ mineralization projects. Besides, Sun et al. (2023) provided a comprehensive review of experimental tools and modeling methods for enhancing the characterization of changes at both the pore and core scales in mineral carbonation, contributing to the advancement of mineralization storage applications. Furthermore, existing reviews in the field of in-situ mineralization often focus on specific types of rocks (Al Kalbani et al. 2023; Rasool and Ahmad 2023). For instance, Oelkers et al. (2018) and Miller et al. (2019c) discussed the primary factors influencing olivine dissolution rates and analyzed the dynamics of its mineralization to better predict and optimize carbonation rate. Raza et al. (2022) delved into the classification and reactivity of basalt minerals, discussing the sequestration potential and emphasized its significance in the context of CO2 storage (Rasool and Ahmad 2023).
Year | Type of mineralization | Key contents and findings | References |
---|---|---|---|
2011 | in-situ | Summarized the natural peridotite carbonation, along with the fundamental principles Proposed methods for in-situ peridotite carbonation | Kelemen et al. (2011) |
2013 | in- & ex-situ | Summarized fundamental processes of carbon mineralization Reviewed the individual capacity and rate of various carbonation strategies | Power et al. (2013) |
2013 | in- & ex-situ | Reviewed the current status of the mineralization sequestration process Summarized the application scenarios of CO2 mineralization | Geerlings and Zevenhoven (2013) |
2014 | in- & ex-situ | Compared the rock types suitable for in-situ mineralization Summarized the processes developed for both rocks and waste resources, along with the reaction kinetics Discussed the utilization of mineralization products | Sanna et al. (2014) |
2015 | ex-situ | Reviewed the current development of pH swing process as one of the most attractive mineral carbonation technologies Relatively high carbonation efficiency was achieved through pH swing mineral carbonation, but the cost was too high | Azdarpour et al. (2015) |
2018 | in- & ex-situ | Introduced the dissolution mechanism of olivine Summarized the factors influencing olivine dissolution rates Listed the applications of olivine dissolution | Oelkers et al. (2018) |
2019 | in- & ex-situ | Compiled a kinetic framework to predict and optimize the carbonation rate of olivine Discussed the factors influencing olivine carbonation | Miller et al. (2019c) |
2020 | in-situ | Introduced the fundamental principles of mineral-carbonation Discussed the potential of mineral carbonation to address the global CCS challenge | Snæbjörnsdóttir et al. (2020) |
2021 | ex-situ | Introduced the pathways and principles for CO2 mineralization Compared the carbonation potential and processes of the most representative and available industrial solid wastes | Liu et al. (2021) |
2022 | in-situ | Discussed different mechanisms and technical challenges of CO2 storage in igneous rocks and proposed a selection criterion based on laboratory and field-scale experience | Raza et al. (2022) |
2023 | in- & ex-situ | Introduced various experimental tools to deepen the characterization of pore-scale and core-scale changes during mineralization storage Summarized multiscale studies of mineralization storage | Sun et al. (2023) |
2023 | in-situ | Reviewed mechanisms and behaviors of the carbon mineralization in basalt, sandstone, carbonate and shale Discussed mainly effect factors of carbon mineralization | Kim et al. (2023) |
2023 | in-situ | Outlined in-situ mineralization in Australia and New Zealand Discussed the different mineralization mechanisms within selected rock types/complexes Examined the governmental funding and perception | Al Kalbani et al. (2023) |
2023 | in-situ | Investigated the reactivity of different basalt minerals Formulated a reactivity scale to assess carbonation potential | Rasool and Ahmad (2023) |
2024 | ex-situ | Summarized CO2 sequestration by industrial wastes through mineral carbonation Indirect carbonation can yield high-purity products | Lin et al. (2024b) |
2024 | in- & ex-situ | Introduced mineral carbonation from the aspects of magnesium-based feedstocks and calcium-based feedstocks The slow kinetics and low carbonation capacity of feedstocks were the main obstacles for industrial application | Lin et al. (2024a) |
2024 | in-situ | Compared the mineralization technologies of the CO2 aqueous solution and wet scCO2 Discussed the similarities and effects of the mineralization influencing factors on the two mineralization methods | Wu et al. (2024) |
Overall, the current state of the field lacks a systematic analysis of the factors influencing the kinetics of in-situ mineralization, and there is also an absence of comprehensive summaries on the dynamic interactions between geochemical reactions and fluid flow. These gaps pose significant limitations to the application of this technology and warrant special attention. Therefore, in this review, we aim to address these gaps by elucidating the feedback at the pore scale during mineralization reactions, discussing the kinetic mechanisms of in-situ mineral carbonation, and providing an overview of the current research status in the field of in-situ mineral carbonation technology. The goal is to explore the limitations of this technology, improve the understanding of in-situ CO2 mineralization, and lay a theoretical foundation for achieving large-scale commercial carbon storage.
In-situ CO2 mineralization primarily relies on the reaction between CO2-rich solutions and mafic/ultramafic rocks. Therefore, the process necessitates the presence of CO2 in a fluid state, which can be done in two ways. The first method involves injecting CO2-saturated solutions (Carbfix approach) (Matter et al. 2011). Its advantage lies in the complete dissolution of CO2 in water, facilitating an immediate reaction with metal cations released from basalt and rapid fixation. From a storage safety perspective, this method virtually eliminates leakage risks, negating the need for an impermeable caprock above the reservoir to act as a barrier. However, a notable drawback is the substantial water usage, requiring at least 10 times the amount to dissolve the gaseous CO2. The second method involves injecting supercritical CO2 liquid (Wallula approach) (Snæbjörnsdóttir et al. 2020), where some of the reservoir pore space is filled with supercritical CO2, referred as residual CO2. Over the span of hundreds to thousands of years, the residual CO2 serves as a continuous source, dissolving into the surrounding water while maintaining a low pH. However, due to the buoyancy of supercritical liquids, ensuring storage safety necessitates the presence of a low-permeability or impermeable caprock both during and after injection, in order to prevent gas leakage before achieving complete mineralization.
The Carbfix project serves as an example to elucidate the mechanisms of in-situ mineralization, and the geochemical process is depicted in Fig. 1 (Sandalow et al. 2021; Noiriel and Daval 2017). The entire process commences with the dissolution of CO2 in water and the transformation of the dissolved inorganic carbon (DIC) species (Eq. (1)). Under specific temperature and pressure conditions, the DIC species are strongly pH-dependent. In the presence of H+, the bridging bonds between cations on the mineral surfaces and silicate groups are disrupted, leading to the release of divalent metal cations (e.g., Ca2+, Mg2+, and Fe2+) into the solution (Eqs. (2)–(4)). Subsequently, these metal ions undergo precipitation reactions with CO32− and HCO3− in the water, resulting in the formation of stable carbonate rock minerals such as calcite (CaCO3), magnesite (MgCO3), and siderite (FeCO3), thereby sequestering free CO2 (Eq. (5)). Meanwhile, there are dynamic interactions between changes in flow and transport properties and geometrical evolution in the system while chemical reactions are taking place. When determining the characteristics of mineral carbonation, it is essential to consider both the time-dependent dynamics and the spatialization of the processes from the atomic scale to the continuum scale. Besides, the diversity of rock minerals in the reaction products is influenced by the types of cations involved, the pH value system and the geological temperature. For instance, under high-temperature conditions (< 300 °C), dissolved calcium tends to readily precipitate calcite or aragonite once the solutions are supersaturated (determined by the ion activity product, IAP, for example [Ca2+] × [CO32−]); at intermediate temperatures (> 65 °C), dissolved magnesium tends to precipitate siderite and magnesite (Snæbjörnsdóttir et al. 2020). Conversely, at low temperatures (< 65 °C), the precipitation of carbonates is kinetically inhibited, resulting in the formation of relatively unstable hydrated carbonate minerals, such as hydromagnesite (Mg5 (CO3)4 (OH)2·4H2O), dypingite (Mg5 (CO3)4 (OH)2·5H2O), and nesquehonite (MgCO3·3H2O) (Turvey et al. 2018). From a thermodynamic perspective, the standard Gibbs free energy of the carbonate products is lower by 0 to 180 kJ/mol compared to that of CO2. This indicates that the formation of carbonate minerals is an energetically favored process, wherein the system transitions from a high-energy state to a low-energy state. As a result, the generated carbonates are relatively stable and less prone to undergoing further decomposition (Snæbjörnsdóttir et al. 2020; Raza et al. 2022; Gras et al. 2020; Gadikota 2021; Turvey et al. 2018).
The rates of mineral carbonation reactions are strongly influenced by the chemical properties and reactivity of different rock types, along with the pressure and temperature during CO2 injection (Kelemen et al. 2019, 2020; Snæbjörnsdóttir et al. 2020). Currently, mafic to ultramafic rock formations predominating in the reservoirs, rich in minerals like wollastonite (CaSiO3), forsterite (Mg2SiO4), antigorite (Mg3Si2O5 (OH)4), and talc (Mg3Si4O10 (OH)2), both on land and in marine environments including continental flood basalts, oceanic basalts, and mantle peridotites, are deemed suitable for in-situ CO2 mineralization. The suitability arises from the abundance of readily soluble alkaline metal ions in these geological structures (Sun et al. 2023; Kelemen et al. 2019). While current research primarily focuses on highly reactive rocks such as basalt, previous studies indicate that carbonation can also occur in other rock types, including siliciclastic reservoirs. However, the carbonation behavior in these reservoirs is intricately linked to mineral composition and heterogeneity (Mishra et al. 2023; Yang et al. 2014; Higgs et al. 2015; Kim et al. 2023). Distinct mineral compositions yield significant variations in geochemical reactions. For instance, the dissolution of minerals like feldspar and chlorite holds substantial potential for enhancing the neutralization of acidic formation water. Ca, Mg, and Fe divalent cations, released by these minerals facilitate the precipitation of carbonate minerals (Mishra et al. 2023). Heterogeneity, on the other hand, introduces spatial variation in geochemical reactions and reservoir properties, serving as impermeable barriers or baffles to CO2 (Yang et al. 2014; Higgs et al. 2015). Although it is commonly believed that the carbonation in low-reactivity rocks like sandstone, may be limited initially, this aspect should not be disregarded, especially when considering the long-term behavior of injected CO2.
Table 2 provides a list of the mineral amount required to sequester one ton of CO2 and the absorption potential for different minerals. It is observed that forsterite is the most effective mineral for optimal sequestration (Raza et al. 2022; Oelkers et al. 2008). However, due to the multitude of parameters influencing mineralization reactions, it is challenging to definitively determine the type of carbonate precipitate and reaction rate during CO2 injection. Therefore, a comprehensive assessment of various factors should be taken into account when evaluating in-situ mineralization sequestration sites.
Minerals | Chemical Formula | Tons required to sequester 1 ton of CO2 | Potential storage rate |
---|---|---|---|
Wollastonite | CaSiO3 | 2.64a | 38% |
Forsterite | Mg2SiO4 | 1.60b | 63% |
Serpentine | Mg3Si2O5 (OH)4 | 2.10b | 48% |
Anorthite | CaAlSi2O8 | 6.30a | 16% |
Basaltic glass | Na0.08K0.08Fe (II)0.17Mg0.28Ca0.26 Al0.36Fe (III nṇṁ)0.02SiTi0.02O3.45 | 2.36c | 42% |
When discussing mafic and ultramafic rocks as candidates for in-situ mineralization, their relative abundance is a crucial consideration. Basalt is abundant and widely distributed, being a major component of the oceanic crust. In contrast, peridotite is quite rare even as an outcrop and is primarily found in the mantle (NASEM 2019; Sanna et al. 2014). This distinction is significant in carbon sequestration and mineralization reactions, as the widespread distribution of basalt makes it more feasible for large-scale applications, whereas the scarcity of peridotite limits its potential for extensive use. Basalt reserves are abundant, constituting a significant portion of the global seafloor (approximately 70% of Earth's surface area) and more than 5% of terrestrial land. It is reported that the carbon storage capacity for onshore basalts globally ranges from approximately 1.0 × 103 ~ 2.5 × 105 Gt (NASEM 2019). While for submarine basalts, it is estimated to be around 8.0 × 102 ~ 1.0 × 105 Gt, along the capacity for onshore peridotites is approximately 6.0 × 104 ~ 6.0 × 105 Gt (NASEM 2019). Therefore, as an effective complement to CCUS technology, in-situ CO2 mineralization can expand the scope of potential storage sites, facilitating better source-sink matching and reducing overall costs. In contrast, the amount of CO2 absorbed through basalt on continents and volcanic islands accounts for 30% of total CO2 absorption through silicate weathering on land, highlighting substantial carbon sequestration potential (Snæbjörnsdóttir et al. 2020). Previous studies (Goldberg and Slagle 2009; Wiese et al. 2008; Snæbjörnsdóttir et al. 2014; Snæbjörnsdóttir and Gislason 2016; McGrail et al. 2006) have conducted rough assessments of carbon sequestration potential in seafloor plateau basalts, mid-ocean ridge basalts, and continental flood basalts. The results indicated that the sequestration capacity of flood basalts near Washington and British Columbia was estimated to be approximately 5.0 × 102 to 2.5 × 103 Gt (Goldberg and Slagle 2009). In the region of Iceland, every 10 m3 of mid-ocean ridge basalt can naturally absorb over 1 metric ton of CO2, leading to an extrapolated theoretical storage potential of 1.0 × 105 to 2.5 × 105 Gt for mid-ocean ridge basalts globally, which surpasses the total amount of CO2 released from fossil fuels (Wiese et al. 2008; Snæbjörnsdóttir et al. 2014; Snæbjörnsdóttir and Gislason 2016). Continental flood basalts are primarily found in the United States and India, with the Columbia River Basalt Group in the United States estimated to have a sequestration potential of approximately 36–148 Gt (McGrail et al. 2006). Peridotite is a key constituent of serpentinite, with the largest known exposure being the Semail serpentinite, approximately 350 km in length, and 40 km in width, with an average thickness of 5 m. Within this formation, about 30% is mantle peridotite (Sanna et al. 2014). Kelemen and Matter (2008) have assessed the immense potential for in-situ mineral carbonation in the Semail serpentinite. They estimated that if all the Mg2+ in the mantle peridotite of the Semail serpentinite were transformed into carbonate minerals, it could sequester 7.0 × 104 Gt of CO2.
However, due to the involvement of complex physicochemical reactions (a complex function of temperature, pressure, wettability, rock properties, and fluid composition) in the in-situ mineralization process, its sequestration potential may change over time (Raza et al. 2022). Consequently, there is no specific calculation method available for accurately assessing the potential of in-situ CO2 mineralization. Current estimation approaches typically relies on initial rock reservoir assessments (Kang et al. 2018; De Silva and Ranjith 2012). These assessments begin with volumetric methods to identify suitable rock reservoirs for in-situ mineralization storage. Subsequently, detailed characterization of mineral components and pore characteristics within the rocks is performed using techniques such as mercury intrusion (MICP) and X-ray computed tomography (X-CT) (Sun et al. 2023). Finally, the storage capacity is inferred based on Eq. 5, while the calculation method for rock reserves is shown in Eq. 6 (Bachu et al. 2007).
where, Vrock is the theoretical volume of the rock, m3; A is the reservoir area, m2; h is the reservoir thickness, m; φ is the reservoir porosity; Sw is the water saturation; Viw and Vpw respectively refer to the volume of water injected into and produced from the reservoir during the recovery or sequestration phase, m3.
Current research often neglects the time-dependent mineralization rates, leading to varied estimation methods and assessments of sequestration potential. Kang et al. (2018) analyzed the distribution of basalt reservoirs in the northern East China Sea through two-dimensional seismic and well logging data, estimating a CO2 storage potential of 0.59 to 2.48 billion metric tons. Based on volumetric methods and mineral trapping mechanisms, Zhang et al. (2023) established an evaluation method for onshore basalt in China, estimating a potential of approximately 4.7 × 104 Gt of CO2 storage, considering direct injection and carbonated water methods. Although the latter had a smaller CO2 storage capacity, its mineralization reaction rate was faster, hence it is preferable. The recent study by Zhang and Tutolo (2022) highlighted the substantial in-situ mineralization potential within glauconitic sandstones in Alberta Province, estimating a capacity to mineralize 150–590 Gt of CO2 based on a reserve assessment of about 1.0 × 103 Gt of glauconite. This notable potential arose from the naturally high proportion (20%–50%) of Fe (II) and Fe (III) in glauconite. Moreover, Ramos et al. (2023) asserted that the bottom flood basalts in the Campos Basin of Brazil exhibited notable CO2 sequestration capabilities. Through modeling of a hypothetical reservoir above the Badejo oilfield, with an area spanning of 31 km2 and a thickness of 300 m, they estimated a storage capacity of CO2 ranging from 16 to 47 Mt. Overall, existing assessments of in-situ mineralization's sequestration capacity remain theoretical, emphasizing the need for detailed, site-specific evaluations to determine effective CO2 storage capacities.
Unlike studies assessing carbon storage potential in sandstone formations, the current evaluation methods for in-situ mineralization sequestration lack a hierarchical and coherent system, similar to the technology-economic resource pyramid proposed by the Carbon Sequestration Leadership Forum (CSLF) (Bachu et al. 2007). While the numerical values for the storage potential presented in the aforementioned studies are promising, the current evaluation methods primarily rely on assumptions of complete conversion of calcium, iron, and magnesium into carbonate minerals (Zhang et al. 2023).Limited effort has been made to establish evaluation methods based on geochemical reaction kinetics and engineering feasibility. As a result, the calculated storage capacity often significantly exceeds the actual effective capacity, typically by orders of magnitude. Furthermore, most existing research overlooks the impact of injectivity on carbon storage, particularly in mafic/ultramafic rock formations with low porosity and permeability, where smooth injection of CO2 may be difficult (He et al. 2024; Kim et al. 2023). Therefore, it is crucial to recognize the fact that existing potential assessments yield theoretical values that are significantly overestimated. To accurately reflect practical in-situ mineralization sequestration capacities, there is a pressing need to develop a more refined evaluation system.
There is a dynamic interplay between chemistry, geometry, flow, and transport processes. Pore-scale feedback between geochemical reactions and fluid flow can induce changes in permeability (see Fig. 1) (Noiriel and Daval 2017). On one hand, as reaction products or secondary minerals accumulate within pore space, they occlude flow path and decrease permeability. Such negative feedback mechanisms can thus influence the flow and transport conditions near mineral interfaces, thereby reducing the mineralization rate (NASEM 2019). This negative feedback is more prone to occur in the carbonation process between CO2-rich fluids and olivine rocks. Peuble et al. (2019) investigated the impact of CO2 and fluid flow on olivine alteration, revealing significant decreases in permeability induced by both carbonate minerals and hydrous minerals. This substantiated the negative feedback effect of olivine alteration products on fluid flow. On the other hand, CO2 injection typically induces rock dissolution, facilitating the opening of conduits, while the volume expansion resulting from carbonate precipitation can create significant stress differentials, potentially leading to the formation of fractures and new fluid pathways. Such a process relies on geometric evolution across various scales, spanning from the crystal surface to the pore space. The differences in reactivity between dissolution along mineral boundaries and reactions between minerals or the formation of secondary precipitates, contribute to altering flow pathways at larger scales. This positive feedback loop may enhance the permeability and expose fresh surfaces to the fluid phase, thereby promoting mineralization reactions (NASEM 2019; Noiriel and Daval 2017; Zhang et al. 2022; Rosenqvist et al. 2023). Further corroboration of this phenomenon was provided by the 3D high-resolution micro-fault scans, revealing that low-temperature carbonation led to conduit occlusion, resulting in reduced permeability (Jöns et al. 2017). Other mineralization experiments have observed the emergence of crack networks due to the precipitation of secondary minerals, leading to an increase in permeability and fluid transport properties (Lisabeth et al. 2017; Kanakiya et al. 2017). These fracture networks were jointly governed by the interaction of physical and chemical stresses, representing an outcome of chemical–mechanical coupling. Concentration gradients existing between secondary and primary pores facilitated the dynamic flow of dissolved material. This flow, coupled with the stress increment in localized secondary pores, culminated in localized strengthening. Meanwhile, the deposition of secondary minerals along the fracture surfaces could laterally inhibit dissolution and deformation, leading localized strengthening through compressive stresses that induce pore bands, thereby aiding in the propagation of near-critical fractures without requiring substantial crystalline pressure from secondary mineral growth within limited pore spaces. Thus, it allows for the gradual growth of fracture networks under near-critical stress intensities (Lisabeth et al. 2017; Kanakiya et al. 2017).
In recent years, there has been increasing attention on the importance of micro- and nano-scale characteristics within fluid-rock systems. These characteristics encompass attributes such as fluid-mineral surface energy, adsorption capacity and separation pressure. Such characteristics play a pivotal role in determining the occurrence of self-limiting negative feedback (occlusion) and the facilitation of accelerating positive feedback (fracturing) (Lambart et al. 2018; Evans et al. 2017). Nevertheless, due to the complex interplay between reaction kinetics and fluid dynamics, accurately predicting mineral dissolution and crystallization (including nucleation and crystal growth) within porous media remains a challenging task (Deng et al. 2022). Menzel et al. (2022) investigated the reasons behind sustained fluid flow during serpentinization mineralization and discovered that the ductile deformation was involved in the process of serpentinite carbonation, primarily driven by grain boundary sliding, associated with particle expansion flow and dissolution–precipitation. Moreover, lithostatic pressure promoted ductile deformation within the reactive medium and generated new pore channels. Peuble et al. (2018) also reached similar conclusions, suggesting that the dissolution–precipitation coupling during olivine carbonation was governed by nanoscale mass transfer mechanisms. These mechanisms maintained a nano-porous interconnected network near the reaction interface, ensuring sufficient permeability and promoting continuous reactions. It is proven that the feedback from external stresses, changes in rheology, and elevated pore pressure parameters enhanced fluid flow and carbonation process, due to the interdependence of pore space alterations and the intricate interplay between transport and reaction kinetics. These factors, ultimately dictate the evolution of rock porosity, permeability, and specific surface area. This relationship can be encapsulated by the Damköhler number (\(D_{{\text{a}}} = \frac{r}{vc}\)), which is defined as the ratio of the reaction rate and the diffusion mass transfer rate at the fluid/solid interface; where, r is the intrinsic reaction rate in mol/m2 s, v is fluid velocity (from Darcy's law), and c is the solubility in pure water of the considered mineral (Osselin et al. 2022; Soulaine et al. 2017). For instance, when the injection rate is too low, it may irreversibly reduce permeability near the injection point, resulting in a lesser degree of mineralization. On the other hand, a higher flow rate generally enhances carbonation, as it sustains higher permeability for longer durations. However, excessively high velocities can lead to surface passivation and diminish the carbonation potential (Andreani et al. 2009; Peuble et al. 2015). Peuble et al. (2015) posited that this arises from the dual control of reaction kinetics and transport processes. At low flow velocities, the elevated concentrations of elements released during silicate dissolution indicated the dominance by chemical kinetics. Conversely, at high flow velocities, there were no changes in the fluid concentration of these elements, indicating that fluids and minerals reached instantaneously local equilibrium and that effective reaction rates were controlled by transport. Thus, it is important to control the injection rate to optimize the in-situ carbonation. Additionally, the injection rate significantly impacts fluid injectivity. At lower CO2 injection rates, continuous injection leads to the evaporation of formation water, causing salt precipitation that can clog pores and reduce reservoir injectivity (He et al. 2022). Increasing the CO2 injection rate can inhibit capillary backflow, effectively mitigating local salt accumulation at the injection point and preventing permeability reduction (Qin et al. 2024; Cui et al. 2023). He et al. (2024) proposed a variable injection strategy, suggesting an early transition from low to high injection rates to prevent solution backflow near the injection interface, thereby minimizing salt accumulation and its adverse effects on injectivity.
In a word, the feedback between reaction transport and fluid flow during in-situ mineralization is highly complex. A thorough and accurate understanding of the physicochemical mechanisms at the fluid–solid interface and the coupled dynamics between geochemical reactions and fluids is challenging. Therefore, developing more precise characterization techniques to elucidate this intricate process is essential.
Understanding the dynamic mechanisms governing mineral dissolution and precipitation is crucial for advancing the kinetics of mineralization reactions. It is proved that the in-situ CO2 mineralization rate depends on factors such as the abundance of reservoir rock minerals and the rate of cation release. This interplay is highly influenced by a constellation of factors, including temperature, pressure, pH, fluid flow velocity, fluid composition, and the interfacial area of the minerals (Zhang and DePaolo 2017; Snæbjörnsdóttir et al. 2020; Matter and Kelemen 2009; Kelemen and Matter 2008).
Due to the complexities of geochemical reactions, there is currently no universally standardized quantitative formula for mineralization kinetics. Depending on the specifics of the reaction conditions, the prevailing approaches often involve the utilization of two dynamic equations, which are Shrinking Core Model (Sun et al. 2008; Wang et al. 2017, 2021a) (Eq. 7) and Avrami Model (Miller et al. 2019c; Wang et al. 2017) (Eq. 8), to calculate the rate of mineralization reactions. However, adjustment should be made for these models by including empirical realities.
where α is the extent of mineralization reaction (%, Eq. (9)); k is the apparent rate constant (min−1); t is the time of mineralization reaction (min); n is an empirical constant depending on the reaction mechanism; \(P_{{\text{CO}_{2} 0}}\) and \(P_{{\text{CO}_{2} t}}\) are the partial pressure of CO2 when the reaction starts and stops, respectively. For the scenario where n = 1, Eq. (7) elucidates a reaction governed by phase boundaries, where the particle surface serves as the reaction zone with rapid and dense heterogeneous nucleation, followed by movement of a reaction front from the surface towards the interior of the particle. When n = 2, Eq. (7) signifies a reaction constrained by diffusion, wherein the rate-determining step involves the diffusion process across a layer of product, and the reaction region diffuses through the product layer into the unreacted core of the particle.
Under the classical homogeneous nucleation theory, nucleation is facilitated through the formation of minute nuclei that traverse the energy barrier within metastable larger-volume pre-existing phases, leading to the emergence of novel phases. This energy barrier, intricately related to interfacial energy and affinity, serves as a pivotal determinant governing both the quantity and attributes of resultant particles (Abdolhosseini Qomi et al. 2022; Wallace et al. 2013). In comparison, heterogeneous nucleation is more advantageous due to its potential to exhibit lower energy barriers (Scheifele et al. 2013). In the initial stage, the formation of a metastable solid phase takes place due to the presence of a smaller nucleation barrier compared to that of inhibiting the nucleation of the stable phase. Eventually, nucleation of the stable phase occurs, either heterogeneously on or within the metastable particles, or homogeneously within the surrounding solution. This event consequently leads to the dissolution or recrystallization of the metastable phase, as commonly observed in systems like calcium carbonate (De Yoreo et al. 2015). Under ambient environmental conditions, calcite is the most stable phase, followed by aragonite, and subsequently vaterite (Declet et al. 2016). The precipitation of CaCO3 initiates from supersaturated solutions, followed by dehydration and recrystallization into vaterite, which is the least stable polymorph. Lastly, vaterite dissolves and undergoes recrystallization into the stable calcite phase (Lin et al. 2022). In the carbonation process of magnesium-rich ores, the formation of magnesite and dolomite (CaMg (CO3)2) within their respective polymorphic/intermediate states is governed by sluggish kinetics, despite the thermodynamic stability. This behavior can be attributed to the following factors: (1) The carbonation process of Mg precursors is predominantly influenced by the formidable energy barrier associated with the dehydration of Mg2+·6H2O cations, leading to a preferential formation of hydrated magnesium carbonate polymorphs. (2) The structural and spatial hindrance presented by the CO32− groups within the rhombohedral arrangement of dolomite and magnesite significantly influences the system, forming the corresponding polymorphs (Santos et al. 2023). As newly formed particles at the mineral–water interface are typically amorphous and hydrated, the physicochemical properties of nuclei and surfaces undergo substantial and dynamic transformations over time and in response to hydration chemistry. To address this challenge, Jun et al. (2016) employed a time-resolved in-situ technique utilizing Small-Angle X-ray Scattering (SAXS) and Grazing Incidence Small-Angle X-ray Scattering (GISAXS) to assess nucleation kinetics, and confirming that by coupling this method with the latest development of surface characterization technology, the understanding of nucleation can be further improved.
Mineral dissolution is the rate-limiting step in in-situ CO2 mineralization sequestration processes, wherein the ability of silicate minerals to release divalent metal cations (e.g., Ca2+, Mg2+, and Fe2+) directly determines the rate of mineralization reactions (Zhang and DePaolo 2017; Daval 2018). Studies have shown that the dissolution rate of wollastonite was the highest, reaching values of 8.0 × 10–9 ~ 2.0 × 10–7 mol/m2·s at 25 °C and 1.6 × 10–5 ~ 5.0 × 10–4 mol/m2·s at 180 °C. However, the availability of wollastonite is limited, with a global estimated reserve of only 100 million tons. In contrast, forsterite exhibited a dissolution rate comparable to that of wollastonite when reacting with CO2-rich fluids. Furthermore, as a principal constituent of Earth's mantle peridotites, forsterite has abundant reserves, and is thus a promising candidate for mineral carbonation sequestration (Kelemen et al. 2019; NASEM 2019). In recent years, a multitude of experiments and numerical investigations have been undertaken to elucidate the kinetics of mineral dissolution and carbonation. However, research specifically centered on in-situ mineralization is still scarce, so we summarized the outcomes of related studies, as documented in Table 3. It is evident that the dissolution rate of minerals is primarily governed by factors such as pH, temperature, and mineral surface area.
Reference | Method | Aqueous Matrix | T (°C) | P (MPa) | pH | Key Remarks | ||
---|---|---|---|---|---|---|---|---|
Miller et al. (2015) | Experimental (High-pressure static reactor) | acetate (C2H3O2–), malonate (C2H3O22–), oxalate (C2O42–), and citrate (C6H5O73–) | 50 | 9 | n/a | Organic ligands enhanced the dehydration of Mg2+, thus influencing the crystallization of nesquehonite and magnesite | ||
Gautier et al. (2016) | Experimental (Mixed-flow reactor) | oxalate, citrate and EDTA | 100, 120, 146 | 1 | < 7 | Organic ligands inhibited magnesite growth through adsorption on the growth-controlling sites at the surface | ||
Xu et al. (2017) | Experimental (Lab-on-a-chip) | Brine | 25 | 9 | n/a | The mineral dissolution and precipitation changed the morphology, porosity, and permeability of the porous rock medium, which then affected the two-phase flow | ||
Farhang et al. (2017) | Experimental (Dissolution set-up) | Buffer solution | 25 | 0.1 | 4.6 | An initial rapid liberation of Mg2+ followed by a very slow extraction was observed, and the extent of extraction depended on the particle size and solution pH | ||
Miller et al. (2018c) | Experimental (In situ X-ray diffraction) | citrate (C6H5O73–) | 50 | 9 | n/a | Citrate promoted the precipitation of magnesite within the nanoscale film due to the partial dehydration of Mg2+ and the adsorption of citrate onto the nucleus and magnesite surfaces | ||
Yadav and Mehra (2019) | Experimental, Mathematical Modelling | Water | 25 ~ 90 | 0.1 | ~ 7 | The carbonation reaction was strongly influenced by the rate of dissolution of wollastonite and the reaction temperature | ||
Kashim et al. (2020) | Experimental (Autoclave reactor) | NaCl brine | 35 ~ 90 | 10 ~ 28 | n/a | The reaction temperature mainly controlled wollastonite in situ mineral carbonation The dissolution mechanism was significantly affected by all factors of pressure, temperature, and salinity | ||
Placencia-Gómez et al. (2020) | Experimental (In situ infrared spectroscopic) | Wet scCO2 | 50 | 9 | n/a | Magnesium carbonate precipitation began at 1.5 monolayers of adsorbed H2O Due to an abrupt decrease in the free-energy barriers for lateral mobility of outer-spherically adsorbed Mg2+ | ||
Marieni et al. (2021) | Numerical Modelling | Seawater | 25–260 | 0.9 | 8.1 | In the seawater systems, the pH was buffered at ≤ 6 for a substantial time due to clay mineral formation, which may lead to somewhat slower carbonation rates | ||
La Plante et al. (2021) | Experimental (Batch reactor) | Pure water | 25, 45, 90 | 0.1 | 4 | The carbonation efficiency was controlled by the atomic topology (network connectivity) of the solid reactant Network rupture was the rate-controlling step of dissolution | ||
Lu et al. (2022) | Experimental (Flow-through dissolution) | H2CO3, HCl, HNO3, H3PO4 | 25 | 0.1 | 4.4 | There was a fast, transient stage, and a slow, stoichiometric stage of dissolution Inorganic ligands, such as sulfate and phosphate, enhanced mineral dissolution rates by forming surface complexes | ||
Oelkers et al. (2022) | Experimental, Numerical Modelling | NaCl brine | 25, 100 | 1 | 3 | Ca and Mg could be preferentially released in the natural system through metal for proton exchange reactions High temperatures favored mineral carbonation | ||
Miao et al. (2023) | Experimental (Batch reactor) | NH4Cl, (NH4)2SO4 and CH3COONH4 | 20–80 | 0.1 | 6–12 | The introduction of ammonium salt resulted in a higher reaction rate and carbonation efficiency The gas–liquid mass transfer was the rate-controlling step | ||
Mesfin et al. (2023) | Experimental (Mixed-flow reactor) | KCl, MgCl2 NaCl, CaCl2 | 25 | 0.1 | 3.6 | The presence of CaCl2 and MgCl2 decreased the labradorite dissolution rates but increased the basaltic glass, due to the increase of ionic strength in aqueous solutions | ||
Delerce et al. (2023) | Experimental (Mixed-flow reactor) | NaCl brine | 120 | 1.5 | 3 | Altered basalts readily released Si, Ca, and Mg to the fluid phase at acidic pH Mineral dissolution rates increased substantially with increasing temperature | ||
Chen et al. (2023) | Experimental | Pure water | 25 | 0.1 | 3–7 | An interfacial reaction with the mutual fusion of ion constituents and replacement of calcium and magnesium occurred between serpentine and calcite during co-milling |
Extensive investigations have shown that the release rates of Ca2+ and Mg2+ from magnesium-iron-bearing minerals during mineral dissolution are significantly influenced by pH. Dissolution rates exhibit a strong dependence on fluid pH, as demonstrated by Snæbjörnsdóttir et al. and illustrated in Fig. 2 (Snæbjörnsdóttir et al. 2020). Notably, aluminum-bearing minerals (such as plagioclase and basaltic glass) exhibit a U-shaped pattern in dissolution rates with increasing pH, with a minimum reaction rate near pH neutral conditions. In contrast, non-aluminous minerals like forsterite and pyroxenes typically experience a continual decline in dissolution rates as pH increases, indicating that these minerals dissolve slowly under conditions conducive to carbonate precipitation (Snæbjörnsdóttir et al. 2020; Gudbrandsson et al. 2011).
a Ca2+ and b Mg2+ release rates from mafic rocks and minerals under different pH at 25 °C. Adapted with permission from Snæbjörnsdóttir et al. (2020)
Low pH conditions are preferable for silicate dissolution, associated with increasing rates of calcium and magnesium leaching. Conversely, higher pH conditions are conducive to carbonate precipitation. Hence, a pH-swing process is engendered (Azdarpour et al. 2015). By adjusting the fluid composition, the acidity of the solution can be changed, thereby facilitating elevated mineralization rates. Research findings indicated that the mechanisms of enhancing mineral dissolution kinetics vary with the kinds of acidic substances added. Acids capable of yielding protonating ions serve to promote dissolution through ligand-mediated pathways, while non-protonating acids (such as HCl and HNO3) expedite dissolution by intensifying overall acidity (Lu et al. 2022; Rashid et al. 2022). Inorganic ligands, including sulfates, carbonates, and phosphates, can enhance mineral dissolution rates through the formation of surface complexes. The dissolution experiments conducted by Lu et al. (2022) revealed that acids containing HCO3− or H2PO4− could adsorb onto the surface of serpentine, forming inorganic ligands. These ligands interacted with the Mg-O bonds and facilitated the detachment of Mg2+ from the surface, ultimately augmenting the initial dissolution rate of serpentine. Organic ligands similarly influenced fluid-rock interactions, thereby regulating mineralization efficiency. Miller et al. (2015) investigated the impact of organic ligands such as acetate, oxalate, citrate, and malonate on mineralization reactions in scCO2 fluid. They discovered that they could enhance the dehydration of Mg2+, thereby facilitating the conversion of hydro-magnesite to magnesite (Miller et al. 2018c). Reynes et al. (2023) employed 2,2'-bipyridine as a ligand to synthesize iron (II) carbonate (FeCO3) through cationic complexation, aiming to facilitate the carbonation process of iron (II)-rich silicate minerals like ferrous olivine. Their results indicated that under conditions of pH = 11 and 80 °C, a maximum carbonation efficiency of 50% could be achieved. However, the use of high organic ligands could hinder the growth of magnesite. The results of Gautier et al. (2016) concluded that the presence of organic ligands inhibited the growth of magnesite. This inhibition primarily resulted from (1) the chelation of Mg2+ in solution, which reduced the solution's saturation state, and (2) the adsorption of organic ligands onto growth sites on the magnesite surface, decreasing the kinetic rate constant of magnesite growth. In pH-swing experiments aimed at regulating alkalinity, most carbonation studies employed buffered solutions of NaHCO3 or KHCO3 with concentrations ranging from 0.4 to 1.0 mol/L, or even as high as 8 mol/L, to maintain pH around 8. However, practical applications in sequestration site aquifers often require consideration of the natural fluid chemistry, which hasn't been extensively utilized for controlling alkalinity (Tutolo et al. 2021).
In recent years, the concept of in-situ pH-swing has been proposed as a means to achieve in-situ mineralization sequestration. Lowering the solution's pH through the introduction of suitable reagents to facilitate the extraction of calcium and magnesium. As the reaction progresses, the pH is subsequently elevated, thereby promoting carbonate precipitation (Azdarpour et al. 2015; Arce et al. 2017). Miao et al. (2023) demonstrated that the introduction of ammonium salt solutions (NH4Cl, (NH4)2SO4, and CH3COONH4) could enhance mineralization reaction rates and carbonate efficiency as a result of increased acidity of ammonium solutions, which facilitated the extraction of Ca2+. Subsequently, the generation of free ammonia shifted the solution to an alkaline state, thus intensifying the kinetics of carbonation. Furthermore, recent investigations by Mesfin et al. (2023) have explored the impact of cationic chloride concentrations on mineral dissolution rates. Their findings revealed that the addition of CaCl2 or MgCl2 enhanced the dissolution rate of basaltic glass while diminishing the dissolution rate of labradorite. This suggested that these divalent metal ions might inhibit the formation of silicon-rich activating complexes on the surface of calcium-rich plagioclase. Similar findings were also reported by Marieni et al. (2021). In the presence of seawater, the precipitation of magnesium-rich aluminum silicates led to pH buffering below 6, resulting in a reduction of mineralization efficiency.
Temperature is another crucial factor influencing mineral dissolution rates, often described by the Arrhenius equation to depict the temperature dependence of reaction rates (Schott et al. 2009). According to predictions, the reaction rate of olivine rock at 30 °C and 0.04 kPa \(P_{{\text{CO}}_2}\) (atmospheric \(P_{{\text{CO}}_2}\)) is approximately a million times slower than at 185 °C and 1.5 × 104 kPa \(P_{{\text{CO}}_2}\) (Kelemen et al. 2011). Research conducted by Oelkers et al. (2022) indicated that the carbonation reaction of basalt was slow at 25 °C. However, at 100 °C, over 95% of the injected dissolved CO2 in the solution would be sequestered as carbonate within five years. (Delerce et al. 2023) investigated the dissolution rates of naturally weathered basalt. Despite cation release rates being 1 to 3 orders of magnitude slower compared to unaltered basalt, mineral dissolution rates significantly increased with rising temperatures. Yadav and Mehra (2019) utilized kinetic models to forecast the potential of in-situ mineralization sequestration for wollastonite. Their findings indicated that the dissolution reaction was governed by the diffusion of various mineral ions through particles, with reaction temperature playing a crucial role in the carbonation process of wollastonite (Kashim et al. 2020). Numerous studies have investigated the dissolution rates of individual common minerals under different pH and temperature conditions. Despite the consistency observed in these findings, there is still ongoing debate regarding their applicability to natural rock-fluid systems. The general expression for mineral dissolution rates is commonly represented as Eq. 3.5 (Zhang and DePaolo 2017). However, specific minerals, such as forsterite (Declercq et al. 2023; Miller et al. 2019c; Rimstidt et al. 2012), basaltic glass (Gudbrandsson et al. 2011; Raza et al. 2022), pyroxenes (Oelkers 2001), and plagioclase (de Obeso et al. 2023; Oelkers 2001) exhibit variations in the calculation process, requiring the optimization of the approach based on experimental data fitting.
where, r is the reaction rate constant; A0 is the temperature-independent pre-exponential factor; Ea is the activation energy; Rmineral is the mineral dissolution rates; Sr is the reaction surface area; k is the absolute kinetic rate constant, and it serves as an explicit function of pH, temperature, or other characteristics of the solution in certain cases; Q is the ionic activity product of the mineral in the solution; K is the equilibrium ionic activity product of the mineral in the solution; α is the empirical constant.
Mechanochemical processes are also considered an important means of influencing mineralization efficiency. Chen et al. (2023) introduced a straightforward method to improve the reactivity and solubility of silicates by co-grinding with calcite, which activates serpentine. During co-grinding, there's an interfacial reaction where mutual ion exchange and calcium-magnesium substitution occur between serpentine and calcite, resulting in the formation of magnesite. Parallel findings were reported by Stillings et al. (2023), who demonstrated that comminution processes augment the CO2 capture proficiency of minerals. This enhancement was attributed to the chemical adsorption of CO2 into the crystalline structures during the fragmentation of rock. The energy liberated upon the rupture of chemical bonds under pressure serves as a heat source, facilitating the requisite energy demand of the reaction.
Furthermore, it is essential to consider both the chemical composition and microstructure of minerals when assessing their carbonation potential. La Plante et al. (2021) compared the carbonation potential of different minerals, revealing that the rate of carbonation and the release of calcium ions were influenced not only by the total Ca (and Mg) content but also by the Si content. This suggested that the dissolution of silicates was controlled by the atomic topology (network connectivity) of the solid reactants. Wang et al. (2021a) conducted an investigation demonstrating that irrespective of the olivine content and mineral composition, the mineral carbonation of natural silicate samples was fundamentally governed by the kinetics of olivine dissolution reactions. Hence, the computation of the effective mineral carbonation potential based on the magnesium and iron content within olivine are of paramount importance for the kinetic analysis of carbonation in natural samples. Van Noort et al. (2013) also highlighted that the increase of less reactive mineral constituents did not inherently constrain the dissolution kinetics of peridotite. Compared with the proportion of highly reactive minerals in ultramafic rocks, the microstructural such as fractures, predominantly influence the rates of peridotite dissolution and carbonation. This is due to the fluid infiltration of fluids along grain boundaries, leading to an amplification in the active surface area of reactive olivine.
It is worth noting that the formation of secondary phases (such as amorphous silica, clays, etc.) at the native mineral-fluid interface in the form of nanoporous layers or alteration layers, as well as the presence of non-reactive phases in heterogeneous rock materials, adds complexity to the prediction of mineral dissolution. This complexity arises due to their influence on material and energy exchange at the fluid-mineral interface (Calabrese et al. 2022). Several studies have indicated that silica-rich passivation layers may form through reprecipitation on minerals, reducing the dissolution rates of rocks and carbonate precipitation by decreasing the surface area of minerals directly exposed to reacting fluids. The extent to which secondary minerals impact the dissolution rate of primary minerals usually depended on factors such as structural compatibility between the two minerals, relative differences in solubility, and the presence of interconnected porous pathways within secondary phases (Oelkers et al. 2018). The molecular dynamics simulations demonstrated by Béarat et al. (2006) revealed that the intrinsic permeability of amorphous SiO2 allowed for the diffusion of reactants through the silica-rich layer, highlighting the dynamic complexity of the passivation layer composition. Daval et al. (2011) indicated that the passivation ability of SiO2 coatings depended on the competition between the intrinsic dissolution rate of silicate and the restructuring (densification) rate of amorphous SiO2. These processes affected the porosity of the passivation layer, thereby regulating its inhibitory effect on the dissolution of the underlying phase (Cailleteau et al. 2008). This also accounted for the observed differences in reactivity between olivine and wollastonite, as wollastonite dissolved more rapidly than olivine. They further noted that the extent of densification of the amorphous SiO2 layer, and consequently its impact on ion transport, could be partially governed by the absolute dissolution rate of the mineral, attributed to the surface-specific rate of densification of the SiO2 layer, which was correlated with the initial dissolution rate of the wollastonite (Daval et al. 2017). Sissmann et al. (2013) contended that passivation was a transient process easily influenced by external forcing, and the mineralogical and passivation characteristics of the interface layer vary with temperature. Similar conclusions were drawn by Johnson et al. (2014), who observed a reduction in the dissolution rate of magnesium olivine at 60 °C due to the formation of an amorphous SiO2 coating on the surface. Yet, the dissolution process did not cease, as the layer became thicker over time. Another crucial factor influencing the passivation effect of the SiO2 layer on the surface of olivine was the presence of trivalent iron (Fe (III)), generated due to the rapid oxidation of Fe2+ released during olivine dissolution. The interaction between Fe (III) and SiO2 under aerobic conditions resulted in an inhibitory effect on olivine dissolution (Sissmann et al. 2013; Saldi et al. 2013, 2015). Further support for the role of Fe in determining secondary coating performance came from studies on Fe-free synthesized magnesite, where the absence of a passivation layer rich in Fe3+-Si prevents reaction inhibition (Li et al. 2018; Miller et al. 2018a, 2019a). Furthermore, the fluid composition played a controlling role in the performance of the passivation layer. NaCl and NaHCO3 have been demonstrated to influence the dissolution and carbonation of magnesium olivine by inhibiting the formation of silica-rich coatings (Miller et al. 2019c).
In-situ CO2 mineralization sequestration follows the classical dissolution–precipitation pathway, leading to the concentration of research efforts on the reactivity of mineral carbonate reactions within CO2-acidified aqueous phases. It is indicated that the mineralization is heavily influenced by water concentration, as water content can impact the dissolution, mass transfer, and carbonate precipitation of silicates (Lin et al. 2022; Thompson et al. 2013). Insufficient water content can drastically lower the leaching rate of cations, then decelerating or even halting the mineralization rate (Wang et al. 2021b). Furthermore, under conditions of limited water content, mineralization reactions can engender a passivation effect, leading to the accumulation of reaction products on the surface of silicate minerals. This phenomenon significantly reduces dissolution and/or carbonation rates (Mergelsberg et al. 2023). The extensive implementation of mineral carbonation sequestration often necessitates the consideration of substantial water consumption. In scenarios involving water-deficient scCO2 injection schemes or reservoirs with less water, scCO2 can dissolve a minor quantity of water, resulting in the formation of a wet scCO2 fluid. This aspect appears to be unfavorable for mineral carbonation. However, recent research has revealed unexpectedly high reactivity between metal silicates and wet scCO2 fluids. A quasi-two-dimensional confined interfacial water film, with a thickness ranging from angstroms to nanometers, forms on the mineral surface, serving as a medium for precipitation and dissolution (Placencia-Gómez et al. 2020; Abdolhosseini Qomi et al. 2022; Kerisit et al. 2012; Miller et al. 2013). As shown in Fig. 3, the crucial steps of mineralization reaction occurring within nanoconfined water films encompass: (1) water separation and nanofilms formation; (2) alterations in interfacial morphology; (3) acceleration of dissolution through H+-promoted and ligand-enhance; (4) formation of interfacial metal-carbonate ion pairs; (5) quasi-2D diffusion of dissolved ions and ion pair; (6) nucleation and growth of the metal carbonate phase (Abdolhosseini Qomi et al. 2022).
The reasons behind the accelerated molecular reaction rates at the water-phase interface remain uncertain, potentially originating from dissolution effects or the intrinsic acid–base characteristics of the water surface (Ruiz-Lopez et al. 2020). When confined at the nanoscale within or between interfaces, its roles as a solvent and a reactant may differ significantly from its behavior in bulk water (Knight et al. 2019, 2020; Cavanaugh et al. 2019). Water adsorbed on the mineral surface adopts a highly ordered state at the molecular scale, restricting its rotational freedom. This gives rise to reduced dielectric constants and diffusion rates, thereby promoting metal complexation and facilitating the formation of stable low-dimensional surface deposits (Knight et al. 2019; Dewan et al. 2013; Fumagalli et al. 2018; Tokunaga et al. 2017). Besides, due to the extensive contact area between the water film and the mineral, the dissolution of minerals leads to a rapid increase in ion concentration, resulting in the relevant solid phase attaining a highly supersaturated state (Wood et al. 2019). On the contrary, mass translocation occurs via lateral permeation within water films, while the vertical mass transport across mineral surfaces encounters substantial hindrance due to the pronounced solubility of adjacent fluids (Placencia-Gómez et al. 2020; Tokunaga et al. 2017). Abdolhosseini Qomi et al. (2022) normalized rate data from prior studies on natural and synthetic forsterite experiments using the Avrami model. They generated graphs correlating reaction temperature and CO2 pressure to compare the difference in mineralization mechanisms and kinetics in bulk water and wet scCO2 fluids. The outcomes, as shown in Fig. 4, revealed higher reactivity of wet scCO2 at lower temperatures and pressures. This heightened reactivity was attributed to the coupled dissolution of nanoscale particle precipitates and micaceous substrates (e.g., KMg3Si3AlO10 (F, OH)2) within nano-constrained water films, resulting in more significant dissolution compared to corresponding bulk liquid experiments.
Carbonation kinetics in water-rich versus wet supercritical CO2 fluids, based on the CO2 pressure–temperature plane. a The triangle and circle markers correspond to synthetic forsterite in wet scCO2 fluid and natural forsterite in high water-to-rock ratio experiments; b–d The marker colors represent (a) time to 50% conversion (t50%), (c) the exponent and (d) the logarithm of precipitation rate constants in first-order Avrami Model fitted to these compiled experiments. Adapted with permission from Abdolhosseini Qomi et al. (2022)
Despite the proven capability of nanoscale water films to enhance mineral carbonation rates, the distinct crystallization pathways of minerals within them remain unclear. Current research endeavors continue to focus on unraveling the stability and kinetic mechanisms of carbonate transformation within nano-confined water films. The investigation by Dasgupta et al. (2023) illuminated that nanoscale confinement significantly diminished the activation barrier for CO2 conversion to H2CO3, and conversely altered the thermochemical effect from bulk aqueous to nano-confined water. Moreover, the presence of charged intermediates became more readily discernible under nanoscale confinement conditions. With the increasing of confinement, the intensified solvation effects and proton transfer enhance the thermodynamic and kinetic attributes of the reaction. Notably, the carbonation mechanism within water films appears to deviate from the conventional understanding that CO2 must primarily dissolve in water. Lee et al. (2015) demonstrated that carbonation reactions preferentially occur at electron-rich terminal Oxygen sites adjacent to cationic vacancies, even with a thin water film. The molecular simulation outcomes of Zare et al. (2022a) indicated that, in comparison to reactions primarily mediated by bulk aqueous environments, silica surfaces in nanoscale water films actively engaged in chemical reactions through long-range hydroxyl transmission and ligand-enhanced dissolution. Within the interfacial water film, reverse proton transfer between bicarbonate and surface hydroxides promoted the formation of carbonate and the surface metal carbonate complexation, consistent with in-situ spectroscopic measurements (Zare et al. 2022b). This difference may affect the macro kinetics of interfacial carbonation. Mergelsberg et al. (2023) observed a distinctive nanoscale passivation effect within the interfacial water film based on the mineralization of basalt with wet scCO2. Amorphous silica exists as a 2 to 3 nm thick magnesium-depleted layer on reacted forsterite particles, resulting in diminished levels of mineral dissolution. The essential understanding of the fundamental mechanisms driving surface passivation during mineralization under water-deficient conditions is crucial for ensuring the rate and efficiency of carbonation. Besides, it was found that there was a threshold for the dependence of mineralization reaction on water, and the kinetics were correlated with the thickness of the water film (Miller et al. 2020; Schaef et al. 2013; Loring et al. 2011). Loring et al. (2015) indicated that mineral carbonation occurs well below the H2O saturation within scCO2. By accounting for water adsorbed on the mineral surface as surface complexes, the average thickness of the water film is only 7–15 Å. Kerisit et al. (2021) proposed that when the thickness of nano-confined water film greater than 1.5 monolayers, it facilitated the growth of magnesite. This was primarily driven by sustained high supersaturation and low H2O activity, which reduced the free energy barrier for lateral migration of Mg2+, enabling the diffusion processes within the water film (Placencia-Gómez et al. 2020). Furthermore, previous research (Miller et al. 2019b, 2018b; Todd Schaef et al. 2013) demonstrated that the reactivity of H2O thin films was linked to their structural arrangement. Specifically, the rates of carbonic acid formation were associated with the concentrations of H2O clusters exhibiting a liquid-like structure.
On the whole, laboratory studies have demonstrated the tendency for reactions to occur in water films. However, visualizing nucleation and growth processes at the nanoscale remains a challenge due to the constraints of the nano-environment. Therefore, gaining a deeper understanding of molecular-scale reaction mechanisms is crucial for enhancing the carbon mineralization efficiency.
To date, a number of geological sequestration demonstration projects have been conducted worldwide, yet only limited real cases for in-situ mineralization projects. The phased progress has only been observed in Iceland, the United States and Japan (Ali et al. 2022; Snæbjörnsdóttir et al. 2020; Lin et al. 2024a). The implementation particulars are comprehensively summarized in Table 4.
Project | Region | Target reservoir | Phase behavior | Operation situation | Mass of injected CO2 | Conversion rate |
---|---|---|---|---|---|---|
CarbFix | Hellisheiði, Iceland | Basalt (400–800 m) | Dissolved CO2 | CarbFix1 referred to the period between 2012 and 2016; CarbFix2 has been in progress since 2014 up to the present day | 101,000 t (Up to Feb.,2024) | 95% |
Wallula | Washington, USA | Flood basalt (800–900 m) | scCO2 | Be initiated in 2009, primarily focusing on subsurface geological characterization; CO2 injection took place from July to August 2013, with operations ceasing in 2015 | 1000 t | 60% |
Nagaoka* | Niigata, Japan | Sandstone aquifer | ||||
(1100 m) | scCO2 | CO2 was injected from 2003 to 2005; Exclusively be utilized for monitoring subsurface activities currently | 10,405 t | n/a | ||
CarbonX | Guangdong, China | Basalt (300–500 m) | Dissolved CO2 | Be scheduled to commence in 2024 | > 1000 t/a | n/a |
The CarbFix project is situated near the Hellisheiði geothermal power plant in southwestern Iceland, which is the third-largest geothermal power station globally. It stands as the largest existing in-situ mineralization sequestration endeavor. Established in 2007, the project officially commenced operations five years later (Matter et al. 2011). Prior to this, carbonate content was measured in drill cuttings from three basaltic geothermal fields, forming the basis for determining the permeability and porosity, and CO2 storage potential of the basalt reservoir within the Icelandic Mid-Atlantic Ridge basalt. The results indicated that every 10 m3 of basalt can naturally sequester over 1 ton of CO2. The target reservoir is located at a depth ranging from 400 to 800 m, characterized by basaltic lithology with minor occurrences of basaltic breccia or intrusive interlayers. The horizontal and vertical permeabilities were measured at 0.3 and 1.7 μm2 respectively. The formation temperature ranged from 30 to 80 °C, with the pH levels between 8.4 and 9.4 (Gislason et al. 2010; Matter et al. 2016). Though the presence of a low-permeability layer overlaying the upper section of the target reservoir, the relatively shallow depth allows for the potential upward migration of CO2 through fractures. Therefore, the design of dissolving CO2 in water before underground injection is considered (see Fig. 5a). In phase 1 (CarbFix1), a total of 230 tons of CO2 were injected, comprising two separate injections of 175 tons of pure CO2 and 73 tons of mixed gas (75 mol% CO2, 24 mol% H2S, and 1 mol% H2) (Sigfusson et al. 2015; Gislason et al. 2010; Galeczka et al. 2022).
Comparison of carbon-injection methods. a The CarbFix method with the dissolution of CO2 in water during injection into a basaltic reservoir and b The Wallula method with liquid CO2 was injected into basalts. Adapted with permission from Snæbjörnsdóttir et al. (2020)
Quantifying the rates of CO2 supply and confirming the presence of precipitated metal phases with identifiable elemental signals pose significant challenges (Gunnarsson et al. 2018). Therefore, monitoring the extent and flux of carbonate and other secondary mineral precipitation solely based on dissolved elemental concentrations is a complex task. it is generally not a common practice to determine carbon sequestration levels by measuring CO2 solubility and analyzing mineral composition post-mineralization reaction during field trials. Instead, the quantity of stored CO2 is typically ascertained through the isotopic tracing method. Tracers employed for in-situ mineralization can be classified into reactive and non-reactive categories. Reactive tracers often involve the use of calcium and magnesium isotopes, which are less influenced by degassing effects (Harrison et al. 2021). Non-reactive tracers frequently utilize carbon isotopes, which effectively differentiate between organic carbon (OC, δ13C values less than 15%) and inorganic carbon (IC, δ13C values greater than 10%). In the case of the CarbFix project, 14C was introduced into the injected CO2 to monitor its transport and reactivity. The amounts of dissolved inorganic carbon (DIC), 14C isotope, and other non-reactive tracers injected were determined, and the difference between these quantities and those measured downstream in monitoring wells was used to calculate the amount of carbon mineralized. Additionally, a mass balance approach was employed to quantitatively assess the fate of the injected CO2. It was observed that nearly all DIC was consumed from injection well to monitoring well. Consequently, it is estimated that over 95% of the injected CO2 had mineralized in less than two years, thereby paving the way for progress into the second phase (CarbFix2) of the project (Matter et al. 2016; Pogge von Strandmann et al. 2019; Snæbjörnsdóttir et al. 2017).
For CarbFix2, the injection rate was significantly increased compared to previous phase. From June 2014 to July 2015, a total of 4526 tons of dissolved CO2 and 2536 tons of dissolved H2S mixed fluid were injected. The target basaltic layer was deeper and at a higher temperature, approximately 250 °C. The monitoring revealed that a significant portion of the injected CO2 began to convert into carbonate minerals within a few months. According to Gunnarsson et al. (2018), 1 ton of CO2 would generate 0.84 m3 of calcite, and 1 ton of H2S would produce 0.70 m3 of pyrite. If all the gases injected at the CarbFix2 site from June 2014 to December 2017 were to precipitate in the form of calcite and pyrite, the total volume of these minerals would be approximately 2.9 × 104 m3. This accounts to only 0.005% of the Hellisheiði reservoir's capacity (which is 6 × 108 m3), highlighting the significant carbon sequestration potential. Additionally, the operating cost of Carbfix is less than 25 US dollars/t CO2, making it competitive with carbon allowance prices in the ETS market and cheaper than other CCS methods. The primary energy requirement for the project is the energy needed to pressurize water containing CO2 to 2.5 MPa at 25 °C, which is approximately 75 kWh (CARBFIX 2022). Currently, this CO2 fixation plant has been operating smoothly, with a total injection of over 80,000 tons of CO2 (CARBFIX 2022).
Although the successful operation of the Carbfix project is encouraging, its success is attributed to specific factors, such as favorable geological conditions, abundant freshwater resources, and well-developed geothermal power plant infrastructure. Therefore, future projects should comprehensively consider the geological structure, mineral composition and structure of the basalt reservoirs, and hydrological characteristics of the target site to ensure feasibility. Furthermore, the CO2 injection method of this project requires a significant amount of water, which may not be suitable for all basalt reservoirs (Marieni et al. 2021). Alternative strategies, such as seawater instead of freshwater, or exploring the mineralization process of wet CO2 fluid described in Sect. 3.3, are effective solutions to this issue, but require further research. Based on this, the project team is also conducting foundational research on dissolving CO2 in seawater before injection, to expand the applicability of this technology in water-scarce regions, coastal, and nearshore areas (CARBFIX 2022). Since September 2021, CarbFix has partnered with Climeworks to conduct experiments involving the annual injection of up to 4,000 tons of CO2 into shallower basalt reservoirs, building on the expansion of direct air capture capabilities. Plans are underway to construct a cross-border carbon transportation and storage hub in Iceland, enabling the maritime transport of CO2 from other regions for injection into the CarbFix site. The goal is to increase the sequestration capacity to 3.0 × 105 tons of CO2 per year, with a gradual expansion to 3 million tons per year by 2030 (CARBFIX 2022). It is estimated that the investment for this project ranges from 220 to 250 million euros, with an estimated annual revenue of approximately 25 to 45 million euros at full operational capacity.
In 2009, the Wallula pilot project was initiated in eastern Washington, USA. It is a part of the U.S. Department of Energy's Regional Carbon Sequestration Partnership Initiative, with an estimated total cost of approximately 2.2 billion US dollars before project implementation. It marked the world's first endeavor to conduct in-situ mineralization sequestration using scCO2 in a flood basalt formation, at a depth of 800 to 900 m (Fig. 5b) (McGrail et al. 2017a, 2014a). The flood basalt formation is composed of multiple successive units of lava flows. A regional aquifer is presented in the distribution area, characterized by ample porosity and lateral connectivity. The internal fluid flow within this formation can extend to kilometer-scale distances (Zakharova et al. 2012; McGrail et al. 2006), rendering it suitable for CO2 injection for sequestration. Furthermore, seismic exploration results indicated the absence of fault structures or fractures in the Wallula region, providing a relatively stable geological condition for storage activities. Above the Columbia River basalt formation, there are impermeable sedimentary layers and basaltic layers that serve as caprocks. These caprocks can impede the flow of CO2, or at the very least, slow down its flow rate, providing ample time for carbonation to occur, and securely immobilizing the CO2 (Zakharova et al. 2012).
The CO2 injection operations for this project were completed within a span of 25 days (from July to August 2013), with an approximate daily injection rate of 40 tons of CO2, totaling 1000 tons injected. Post-injection well analyses revealed that the CO2 was located at the top of the basalt flow layer and did not exhibit any upward movement. Additionally, testing of the shallow soil gases around the injection well detected no leaks, indicating a successful sequestration process. After two years of continuous monitoring following the injection, researchers retrieved cores from the injection zone and subjected them to detailed physical and chemical analyses. Nodules found within the entirety of the cores were identified as the carbonate mineral siderite, containing calcium, iron, magnesium, and manganese. Further carbon isotope analysis revealed that these nodules chemically differed from naturally occurring carbonates in the basalt, yet exhibited a pronounced correlation with the isotopic signature of the injected CO2, providing additional evidence of carbonation reactions occurring with the injected CO2 (McGrail et al. 2014b, 2017b). White et al. (2020) conducted an analysis based on hydrological modeling to assess the changes in near-field, wellbore, and reservoir conditions before and after CO2 injection in this project. Their findings indicated that approximately 60% of the injected CO2 was sequestered through mineralization within two years, with the generated carbonates occupying around 4% of the available reservoir pore space. Additionally, Polites et al. (2022) observed the precipitation of calcite and amorphous silica in sidewall cores. They pointed out the co-occurrence of carbonates after CO2 injection, emphasizing the crucial significance of clarifying changes in carbonate nucleation composition for predicting carbonate stability. This insight aided in estimating the rates, timing, and reaction pathway relationships of mineral dissolution and precipitation.
Compared to Carbfix, Wallula project primarily employs conventional post-injection monitoring methods, including in-well monitoring and sampling analysis. During the injection and sequestration process, subsurface temperature and pressure are closely monitored to promptly understand the injection conditions, and fluid samples are collected for hydrological analysis. However, the monitoring data obtained from these methods cannot estimate the percentage of injected CO2 that has undergone carbonation. To achieve better monitoring results, more innovative monitoring methods should be proposed in combination with practical conditions.
The Nagaoka project represents Japan's inaugural carbon sequestration pilot initiative, located in the Nagaoka oil and gas field in southern Niigata, and the well configuration at the reservoir depth is shown in Fig. 6a. The target reservoir is situated at a depth of approximately 1,100 m with a thickness of 60 m. The formation exhibits a subsurface water temperature of 48 °C and a pressure of 10.8 MPa, while the injected CO2 is in a supercritical state. Based on well testing and logging data, the reservoir has been classified into five zones, with Zone 2 (with a thickness of 12 m and porosity of 22.5%) serving as the primary CO2 sequestration layer. The caprock consists of non-permeable Pleistocene shallow marine formations (White et al. 2006; Mito et al. 2008).
a Plane view of well locations at the top layer of Zone 2 and b Reservoir simulation results for CO2 storage in the three phases. Adapted with permission from Mito et al. (2008)
Between July 2003 and January 2005, approximately 20 to 40 tons of high-concentration CO2 (99.9%) were injected daily into the target Haizume formation. Over 17 months, the total CO2 injection reached 10,400 tons, with a sequestration cost of approximately 44 to 95 US dollars/t CO2. The project concluded in 2007, with subsequent efforts primarily focused on reservoir monitoring. Mito et al. (2008) demonstrated through reservoir modeling that mineral trapping likely occurred in the early stages of CO2 sequestration, as shown in Fig. 6b. Additionally, a geochemical TOUGHREACT model established by Yamamoto et al. (2017) indicated that due to pressure-driven processes during injection, reservoir water was displaced, facilitating the gradual diffusion of CO2. Then it promoted the dissolution of rock minerals (such as plagioclase and chlorite) within the Haizume formation, leading to the generation of more stable carbonate precipitates. It also provided conclusive evidence for the existence of the mineralization mechanism. Building on these findings, Japan officially initiated its CCS demonstration project in 2008, conducting additional geological and engineering assessments.
For the key parameter of CO2 saturation, which is crucial for predicting CO2 sequestration amount, the challenge lies in accurately obtaining the CO2 distribution at well points, particularly after injection, as measuring CO2 saturation in cased wells presents significant difficulties. The Nagaoka project in Japan was the first in the world to conduct more than 40 in-well geophysical logging sessions after CO2 injection, setting a precedent for using multiple logging sessions to identify CO2 changes before and after injection and at different stages (Ma et al. 2022a). The execution of this project demonstrated the feasibility of in-situ mineralization in siliceous reservoirs. However, appropriate porosity and permeability are prerequisites for CO2 sequestration, as low permeability may result in low storage capacity and hinder CO2 injection (Kirmani et al. 2024). Strictly speaking, this project does not fall under the category of purely in-situ mineralization sequestration but rather constitutes an onshore saline aquifer storage initiative within siliciclastic reservoirs. However, subsequent advancements in the project have confirmed the presence of in-situ mineral carbonation processes. Therefore, the successful implementation of the aforementioned projects serves as a valuable reference for the advancement of in-situ mineralization.
In 2023, Tencent launched the CarbonX Program, aiming to support the technological development of CCUS in China, by identifying the next generation of cutting-edge low-carbon technologies (Tencent 2023). A total of 30 leading CCUS projects receiving the fund, anticipated to scale up by 2030. One of these projects focuses on the research and application of rapid CO2 mineralization in basalt formations. In addition, Tencent collaborated with Carbfix to establish the first demonstration project for CO2 in-situ mineralization in Asia since 2022. Initial stages such as preliminary research and site selection have been completed, and Leizhou Peninsula of Guangdong Province, China is selected as the sequestration site for further investigation after comparing the distribution and reservoir properties of basalt formations. The basalt formations in the Leizhou Peninsula were formed by mantle plume volcanic activity over millions of years. These formations encompass a vast area (approximately 3,000 square kilometers), primarily composed of alkali olivine and tholeiite. They have undergone low degrees of alteration, exhibit good porosity and permeability, and possess a certain potential for mineral carbonation (Li et al. 2023). The target reservoir is situated at a depth of approximately 300 to 500 m, with suitable porosity and permeability, making it conducive to CO2 injection. Li et al. (2023) have calculated the theoretical storage potential based on the mechanisms of volcanic rock mineralization, taking factors such as the distribution area, average thickness, and mineral composition of the basalt formations into consideration. The results indicated that the total volume of basalt formations on the Leizhou Peninsula is approximately 257 km3. The theoretical sequestration potential of CO2 in this region falls between 19 to 459 billion tons. Furthermore, Carbfix is concurrently developing a containerized injection system design, which is scheduled to commence in 2024, involving the injection and sequestration of thousands of tons of CO2 as part of the carbon fixation demonstration (CARBFIX 2022).
To sup up, in-situ CO2 mineralization sequestration offers the following advantages in reducing CO2 emission: (1) The naturally widespread minerals are suitable for in-situ mineralization, presenting considerable potential for CO2 storage. (2) The geochemical reactions occur rapidly, enabling the conversion of most CO2 into solid carbonate minerals within a short period. This ensures rapid, permanent, and safe CO2 sequestration, reducing reliance on reservoir cap conditions. (3) The injected CO2 can achieve a stable mineral sequestration state relatively faster, avoiding the need for long-term safety monitoring of CO2 sequestration, thereby reducing sequestration costs. The estimated cost for CO2 storge in basalt ranges from 20 to 30 US dollars per ton of CO2.
Nevertheless, several challenges need to be addressed:
Although the mechanism of in-situ mineralization reactions has been extensively studied, the relationship between geochemical reaction characteristics and storage rates and efficiency remains unclear. It is essential to focus on the effects of various factors on mineralization outcomes, clarifying the formation mechanisms of secondary mineral diversity, and revealing the chemical kinetics of mineral dissolution and reprecipitation at different scales to further enhance mineralization rates and efficiency.
Current evaluation methods for in-situ mineralization sequestration potential are relatively simplistic and approximate, often overestimating the actual effective capacity. Establishing a comprehensive and reliable system for selecting target reservoirs and evaluating CO2 storage capacity, considering CO2 injectability, is imperative.
There are two existing forms of CO2 injection: saturated solution and supercritical fluid, each with distinct advantages and disadvantages. Detailed design of CO2 injection schemes, along with technical, economic, and leakage risk assessments, tailored to different types of target reservoirs, is necessary to identify the most suitable implementation plan.
Source-sink matching is a critical for CCUS projects. Matching CO2 emission sources and suitable reservoirs may be challenged by long distances and high costs of pipeline infrastructure. When integrated with direct air capture (DAC), this technology becomes a negative carbon technology, which is beneficial for carbon dioxide removal (CDR). This approach not only effectively removes CO2 from the atmosphere but also permanently sequesters it through mineralization reactions, thereby significantly enhancing the sustainability and efficacy of CCUS projects.
CO2 in-situ mineralization sequestration involves complex multi-process, multi-spatial, and long-term procedures. The safety of storage and its potential negative environmental impacts remain significant public concerns. Therefore, it is crucial to conduct safety and environmental monitoring and impact assessments, ensuring real-time detection of CO2 leakage and timely adjustments to ensure project safety.
In-situ mineralization sequestration technology aims to permanently sequester CO2 from emission sources into reactive rocks, avoiding the steps and costs associated with large-scale mining, transportation, and handling of solid reactants. When integrated with DAC technology, it can play a significant role in achieving carbon neutrality. However, due to the complexity of interactions between reaction transport and fluid flow, accurately measuring the reaction kinetics of carbon mineralization is challenging. Mineral dissolution is the rate-limiting step in the in-situ CO2 mineralization, where the rate of mineralization reactions is directly influenced by the capacity of silicate minerals to release divalent metal cations. This dependence correlates closely with factors such as temperature, pressure, pH levels, fluid composition, and the contact surface area of the minerals. Moreover, water content significantly impacts the dissolution and mass transfer of silicate minerals, as well as the precipitation of carbonate minerals. Recent studies have also emphasized the significance of molecular-scale mechanisms in CO2 mineralization within nanoscale interfacial water films, highlighting the critical nature of these interactions.
The potential for in-situ mineralization sequestration is immense. Undeniably, accurately controlling the optimal conditions and key parameters for mineralization reactions presents significant challenges. Enhancing the rate and efficiency of mineralization reactions remains a primary focus for future work. The low porosity and permeability of mafic and ultramafic rocks pose considerable challenges for CO2 injection and sequestration, thereby limiting the effectiveness of in-situ mineralization. Reliable assessment methods for storage capacity, considering reservoir injectability, need to be developed. The approach of CO2 injection should align with the geological conditions of the sequestration site. Supercritical CO2 can be injected if the reservoir is sufficiently deep with excellent caprock conditions. Otherwise, CO2-saturated solutions are preferable for shallower depths with fractures or faults. In addition to conventional physico-chemical detection methods, further development of precise numerical simulations is required to deepen the understanding of multi-scale fluid-rock reaction kinetics and migration patterns, thereby enhancing visual observation and monitoring techniques.
[1] | Abdolhosseini Qomi MJ, Miller QR, Zare S, Schaef HT, Kaszuba JP, Rosso KM (2022) Molecular-scale mechanisms of CO2 mineralization in nanoscale interfacial water films. Nat Rev Chem 6(9):598–613 |
[2] | Al Kalbani M, Serati M, Hofmann H, Bore T (2023) A comprehensive review of enhanced in-situ CO2 mineralisation in Australia and New Zealand. Int J Coal Geol 276:104316. https://doi.org/10.1016/j.coal.2023.104316 |
[3] | Ali M, Jha NK, Pal N, Keshavarz A, Hoteit H, Sarmadivaleh M (2022) Recent advances in carbon dioxide geological storage, experimental procedures, influencing parameters, and future outlook. Earth Sci Rev 225:103895. https://doi.org/10.1016/j.earscirev.2021.103895 |
[4] | Aminu MD, Nabavi SA, Rochelle CA, Manovic V (2017) A review of developments in carbon dioxide storage. Appl Energy 208:1389–1419. https://doi.org/10.1016/j.apenergy.2017.09.015 |
[5] | Andreani M, Luquot L, Gouze P, Godard M, Hoisé E, Gibert B (2009) Experimental study of carbon sequestration reactions controlled by the percolation of CO2-rich brine through peridotites. Environ Sci Technol 43(4):1226–1231. https://doi.org/10.1021/es8018429 |
[6] | Arce GLAF, Soares Neto TG, Ávila I, Luna CMR, Carvalho JA (2017) Leaching optimization of mining wastes with lizardite and brucite contents for use in indirect mineral carbonation through the pH swing method. J Clean Prod 141:1324–1336. https://doi.org/10.1016/j.jclepro.2016.09.204 |
[7] | Azdarpour A, Asadullah M, Mohammadian E, Hamidi H, Junin R, Karaei MA (2015) A review on carbon dioxide mineral carbonation through pH-swing process. Chem Eng J 279:615–630. https://doi.org/10.1016/j.cej.2015.05.064 |
[8] | Bachu S, Bonijoly D, Bradshaw J, Burruss R, Holloway S, Christensen NP, Mathiassen OM (2007) CO2 storage capacity estimation: methodology and gaps. Int J Greenhouse Gas Control 1(4):430–443 |
[9] | Béarat H, McKelvy MJ, Chizmeshya AVG, Gormley D, Nunez R, Carpenter RW, Squires K, Wolf GH (2006) Carbon sequestration via aqueous olivine mineral carbonation: role of passivating layer formation. Environ Sci Technol 40(15):4802–4808. https://doi.org/10.1021/es0523340 |
[10] | Benson SM, Cole DR (2008) CO2 sequestration in deep sedimentary formations. Elements 4(5):325–331 |
[11] | Cailleteau C, Angeli F, Devreux F, Gin S, Jestin J, Jollivet P, Spalla O (2008) Insight into silicate-glass corrosion mechanisms. Nat Mater 7(12):978–983 |
[12] | Calabrese S, Wild B, Bertagni MB, Bourg IC, White C, Aburto F, Cipolla G, Noto LV, Porporato A (2022) Nano-to global-scale uncertainties in terrestrial enhanced weathering. Environ Sci Technol 56(22):15261–15272 |
[13] | CARBFIX (2022) 'Permanent and secure Geological storage of CO2 by in-situ carbon mineralization'. Iceland |
[14] | Cavanaugh J, Whittaker ML, Joester D (2019) Crystallization kinetics of amorphous calcium carbonate in confinement. Chem Sci 10(19):5039–5043 |
[15] | Chen M, Zhang Q, Li Z, Hu H, Wang C (2023) Insights into the mechanochemical interfacial interaction between calcite and serpentine: implications for ambient CO2 capture. J Clean Prod 401:136715 |
[16] | Cui G, Hu Z, Ning F, Jiang S, Wang R (2023) A review of salt precipitation during CO2 injection into saline aquifers and its potential impact on carbon sequestration projects in China. Fuel 334:126615. https://doi.org/10.1016/j.fuel.2022.126615 |
[17] | Dasgupta N, Ho TA, Rempe SB, Wang Y (2023) Hydrophobic nanoconfinement enhances CO2 conversion to H2CO3. J Phys Chem Lett 14(6):1693–1701. https://doi.org/10.1021/acs.jpclett.3c00124 |
[18] | Daval D (2018) Carbon dioxide sequestration through silicate degradation and carbon mineralisation: promises and uncertainties. Npj Mater Ation 2(1):11. https://doi.org/10.1038/s41529-018-0035-4 |
[19] | Daval D, Sissmann O, Menguy N, Saldi GD, Guyot F, Martinez I, Corvisier J, Garcia B, Machouk I, Knauss KG, Hellmann R (2011) Influence of amorphous silica layer formation on the dissolution rate of olivine at 90°C and elevated pCO2. Chem Geol 284(1):193–209. https://doi.org/10.1016/j.chemgeo.2011.02.021 |
[20] | Daval D, Bernard S, Rémusat L, Wild B, Guyot F, Micha JS, Rieutord F, Magnin V, Fernandez-Martinez A (2017) Dynamics of altered surface layer formation on dissolving silicates. Geochim Cosmochim Acta 209:51–69. https://doi.org/10.1016/j.gca.2017.04.010 |
[21] | de Obeso JC, Awolayo AN, Nightingale MJ, Tan C, Tutolo BM (2023) Experimental study on plagioclase dissolution rates at conditions relevant to mineral carbonation of seafloor basalts. Chem Geol 620:121348 |
[22] | De Silva PNK, Ranjith P (2012) A study of methodologies for CO2 storage capacity estimation of saline aquifers. Fuel 93:13–27 |
[23] | De Yoreo JJ, Gilbert PU, Sommerdijk NA, Penn RL, Whitelam S, Joester D, Zhang H, Rimer JD, Navrotsky A, Banfield JF (2015) Crystallization by particle attachment in synthetic, biogenic, and geologic environments. Science 349(6247):aaa6760 |
[24] | Declercq J, Bowell R, Brough C, Barnes A, Griffiths R (2023) Role of Mg gangue minerals in natural analogue CO2 sequestration. Econ Geol 118(3):675–688 |
[25] | Declet A, Reyes E, Suárez O (2016) Calcium carbonate precipitation: a review of the carbonate crystallization process and applications in bioinspired composites. Rev Adv Mater Sci. 44 (1) |
[26] | Delerce S, Bénézeth P, Schott J, Oelkers EH (2023) The dissolution rates of naturally altered basalts at pH 3 and 120°C: Implications for the in-situ mineralization of CO2 injected into the subsurface. Chem Geol 621:121353 |
[27] | Deng H, Poonoosamy J, Molins S (2022) A reactive transport modeling perspective on the dynamics of interface-coupled dissolution-precipitation. Appl Geochem 137:105207. https://doi.org/10.1016/j.apgeochem.2022.105207 |
[28] | Dewan S, Yeganeh MS, Borguet E (2013) Experimental correlation between interfacial water structure and mineral reactivity. J Phys Chem Lett 4(11):1977–1982 |
[29] | Evans O, Spiegelman MW, Kelemen PB (2017) 'Reactive-brittle dynamics in peridotite alteration' AGU Fall Meeting Abstracts. pp. DI51B-0316 |
[30] | Farhang F, Rayson M, Brent G, Hodgins T, Stockenhuber M, Kennedy E (2017) Insights into the dissolution kinetics of thermally activated serpentine for CO2 sequestration. Chem Eng J 330:1174–1186 |
[31] | Fumagalli L, Esfandiar A, Fabregas R, Hu S, Ares P, Janardanan A, Yang Q, Radha B, Taniguchi T, Watanabe K (2018) Anomalously low dielectric constant of confined water. Science 360(6395):1339–1342 |
[32] | Gadikota G (2021) Carbon mineralization pathways for carbon capture, storage and utilization. Commun Chem 4(1):23 |
[33] | Galeczka IM, Stefánsson A, Kleine BI, Gunnarsson-Robin J, Snæbjörnsdóttir SÓ, Sigfússon B, Gunnarsdóttir SH, Weisenberger TB, Oelkers EH (2022) A pre-injection assessment of CO2 and H2S mineralization reactions at the Nesjavellir (Iceland) geothermal storage site. Int J Greenhouse Gas Control 115:103610. https://doi.org/10.1016/j.ijggc.2022.103610 |
[34] | Gautier Q, Bénézeth P, Schott J (2016) Magnesite growth inhibition by organic ligands: An experimental study at 100, 120 and 146°C. Geochim Cosmochim Acta 181:101–125. https://doi.org/10.1016/j.gca.2016.02.028 |
[35] | GCCSI (2023) Global Status of CCS 2023. Global CCS Institute, Melbourne |
[36] | Geerlings H, Zevenhoven R (2013) CO2 mineralization—bridge between storage and utilization of CO2. Annu Rev Chem Biomol Eng 4:103–117 |
[37] | Gislason SR, Wolff-Boenisch D, Stefansson A, Oelkers EH, Gunnlaugsson E, Sigurdardottir H, Sigfusson B, Broecker WS, Matter JM, Stute M (2010) Mineral sequestration of carbon dioxide in basalt: A pre-injection overview of the CarbFix project. Int J Greenhouse Gas Control 4(3):537–545 |
[38] | Goldberg D, Slagle AL (2009) A global assessment of deep-sea basalt sites for carbon sequestration. Energy Procedia 1(1):3675–3682 |
[39] | Gras A, Beaudoin G, Molson J, Plante B (2020) Atmospheric carbon sequestration in ultramafic mining residues and impacts on leachate water chemistry at the Dumont Nickel Project, Quebec. Canada Chemical Geology 546:119661 |
[40] | Gudbrandsson S, Wolff-Boenisch D, Gislason SR, Oelkers EH (2011) An experimental study of crystalline basalt dissolution from 2⩽pH⩽11 and temperatures from 5 to 75 °C. Geochim Cosmochim Acta 75(19):5496–5509 |
[41] | Gunnarsson I, Aradóttir ES, Oelkers EH, Clark DE, Arnarson MÞ, Sigfússon B, Snæbjörnsdóttir SÓ, Matter JM, Stute M, Júlíusson BM (2018) The rapid and cost-effective capture and subsurface mineral storage of carbon and sulfur at the CarbFix2 site. Int J Greenhouse Gas Control 79:117–126 |
[42] | Harrison AL, Bénézeth P, Schott J, Oelkers EH, Mavromatis V (2021) Magnesium and carbon isotope fractionation during hydrated Mg-carbonate mineral phase transformations. Geochim Cosmochim Acta 293:507–524. https://doi.org/10.1016/j.gca.2020.10.028 |
[43] | He D, Xu R, Ji T, Jiang P (2022) Experimental investigation of the mechanism of salt precipitation in the fracture during CO2 geological sequestration. Int J Greenhouse Gas Control 118:103693. https://doi.org/10.1016/j.ijggc.2022.103693 |
[44] | He D, Wang Z, Yuan H, Zhang M, Hong Z, Xu R, Jiang P, Chen S (2024) Experimental investigation of salt precipitation behavior and its impact on injectivity under variable injection operating conditions. Gas Sci Eng 121:205198. https://doi.org/10.1016/j.jgsce.2023.205198 |
[45] | Hepburn C, Adlen E, Beddington J, Carter EA, Fuss S, Mac Dowell N, Minx JC, Smith P, Williams CK (2019) The technological and economic prospects for CO2 utilization and removal. Nature 575(7781):87–97. https://doi.org/10.1038/s41586-019-1681-6 |
[46] | Higgs KE, Haese RR, Golding SD, Schacht U, Watson MN (2015) The Pretty Hill Formation as a natural analogue for CO2 storage: An investigation of mineralogical and isotopic changes associated with sandstones exposed to low, intermediate and high CO2 concentrations over geological time. Chem Geol 399:36–64. https://doi.org/10.1016/j.chemgeo.2014.10.019 |
[47] | IEA License: CC BY 4.0 (2021) ' Net Zero by 2050'. Paris IEA. Available at: https://www.iea.org/reports/net-zero-by-2050. |
[48] | IEA License: CC BY 4.0 (2024) 'CO2 Emissions in 2023'. Paris: IEA. Available at: https://www.iea.org/reports/co2-emissions-in-2022. |
[49] | Jayasekara D, Ranjith P, Wanniarachchi W, Rathnaweera T (2020) Understanding the chemico-mineralogical changes of caprock sealing in deep saline CO2 sequestration environments: a review study. J Supercrit Fluids 161:104819 |
[50] | Johnson NC, Thomas B, Maher K, Rosenbauer RJ, Bird D, Brown GE (2014) Olivine dissolution and carbonation under conditions relevant for in situ carbon storage. Chem Geol 373:93–105. https://doi.org/10.1016/j.chemgeo.2014.02.026 |
[51] | Jöns N, Kahl W-A, Bach W (2017) Reaction-induced porosity and onset of low-temperature carbonation in abyssal peridotites: Insights from 3D high-resolution microtomography. Lithos 268–271:274–284. https://doi.org/10.1016/j.lithos.2016.11.014 |
[52] | Jun Y-S, Kim D, Neil CW (2016) Heterogeneous nucleation and growth of nanoparticles at environmental interfaces. Acc Chem Res 49(9):1681–1690. https://doi.org/10.1021/acs.accounts.6b00208 |
[53] | Kanakiya S, Adam L, Esteban L, Rowe MC, Shane P (2017) Dissolution and secondary mineral precipitation in basalts due to reactions with carbonic acid. J Geophys Res Solid Earth 122(6):4312–4327 |
[54] | Kang M, Kim JH, Kim KO, Cheong S, Shinn YJ (2018) 'Assessment of CO2 storage capacity for basalt caprock-sandstone reservoir system in the northern East China Sea' EGU General Assembly Conference Abstracts. p. 5772 |
[55] | Kashim MZ, Tsegab H, Rahmani O, Abu Bakar ZA, Aminpour SM (2020) Reaction mechanism of wollastonite in situ mineral carbonation for CO2 sequestration: effects of saline conditions, temperature, and pressure. ACS Omega 5(45):28942–28954. https://doi.org/10.1021/acsomega.0c02358 |
[56] | Kelemen PB, Matter J (2008) In situ carbonation of peridotite for CO2 storage. Proc Natl Acad Sci 105(45):17295–17300 |
[57] | Kelemen PB, Matter J, Streit EE, Rudge JF, Curry WB, Blusztajn J (2011) Rates and mechanisms of mineral carbonation in peridotite: natural processes and recipes for enhanced, in situ CO2 capture and storage. Annu Rev Earth Planet Sci 39:545–576 |
[58] | Kelemen P, Benson SM, Pilorgé H, Psarras P, Wilcox J (2019) An overview of the status and challenges of CO2 storage in minerals and geological formations. Front Clim 1:9 |
[59] | Kelemen PB, McQueen N, Wilcox J, Renforth P, Dipple G, Vankeuren AP (2020) Engineered carbon mineralization in ultramafic rocks for CO2 removal from air: Review and new insights. Chem Geol 550:119628. https://doi.org/10.1016/j.chemgeo.2020.119628 |
[60] | Kerisit S, Weare JH, Felmy AR (2012) Structure and dynamics of forsterite–scCO2/H2O interfaces as a function of water content. Geochim Cosmochim Acta 84:137–151. https://doi.org/10.1016/j.gca.2012.01.038 |
[61] | Kerisit SN, Mergelsberg ST, Thompson CJ, White SK, Loring JS (2021) Thin water films enable low-temperature magnesite growth under conditions relevant to geologic carbon sequestration. Environ Sci Technol 55(18):12539–12548. https://doi.org/10.1021/acs.est.1c03370 |
[62] | Kim K, Kim D, Na Y, Song Y, Wang J (2023) A review of carbon mineralization mechanism during geological CO2 storage. Heliyon 9(12):e23135. https://doi.org/10.1016/j.heliyon.2023.e23135 |
[63] | Kirmani FUD, Raza A, Ahmad S, Arif M, Mahmoud M (2024) A holistic overview of the in-situ and ex-situ carbon mineralization: Methods, mechanisms, and technical challenges. Sci Total Environ 943:173836. https://doi.org/10.1016/j.scitotenv.2024.173836 |
[64] | Knight AW, Kalugin NG, Coker E, Ilgen AG (2019) Water properties under nano-scale confinement. Sci Rep 9(1):8246 |
[65] | Knight AW, Ilani-Kashkouli P, Harvey JA, Greathouse JA, Ho TA, Kabengi N, Ilgen AG (2020) Interfacial reactions of Cu (II) adsorption and hydrolysis driven by nano-scale confinement. Environ Sci Nano 7(1):68–80 |
[66] | La Plante EC, Mehdipour I, Shortt I, Yang K, Simonetti D, Bauchy M, Sant GN (2021) Controls on CO2 mineralization using natural and industrial alkaline solids under ambient conditions. ACS Sustain Chem Eng 9(32):10727–10739 |
[67] | Lambart S, Savage H, Robinson B, Kelemen P (2018) Experimental investigation of the pressure of crystallization of Ca (OH) 2: Implications for the reactive cracking process. Geochem Geophys Geosyst 19(9):3448–3458 |
[68] | Lee M-S, Peter McGrail B, Rousseau R, Glezakou V-A (2015) Structure, dynamics and stability of water/scCO2/mineral interfaces from ab initio molecular dynamics simulations. Sci Rep 5(1):14857. https://doi.org/10.1038/srep14857 |
[69] | Li Z, Guo J, Dong Z, Chen J (2018) Insight into interactions of olivine-scCO2-water system at 140 °C and 15 MPa during CO2 mineral sequestration. Geosci Front 9(6):1945–1955. https://doi.org/10.1016/j.gsf.2017.12.008 |
[70] | Li P, Jiang J, Cheng J, Zhao M (2023) Assessment of carbon dioxide mineralization sequestration potential of volcanic rocks in Leizhou Peninsula, Guangdong Province China. Geol J China Ersities 29(01):76–84. https://doi.org/10.16108/j.issn1006-7493.2022078 |
[71] | Li L, Liu Q, Huang T, Peng W (2022) Mineralization and utilization of CO2 in construction and demolition wastes recycling for building materials: a systematic review of recycled concrete aggregate and recycled hardened cement powder. Separation and Purification Technology 121512 |
[72] | Lin J-Y, Garcia EA, Ballesteros FC, Garcia-Segura S, Lu M-C (2022) A review on chemical precipitation in carbon capture, utilization and storage. Sustain Environ Res 32(1):1–20 |
[73] | Lin X, Li X, Liu H, Boczkaj G, Cao Y, Wang C (2024a) A review on carbon storage via mineral carbonation: Bibliometric analysis, research advances, challenges, and perspectives. Sep Purif Technol 338:126558. https://doi.org/10.1016/j.seppur.2024.126558 |
[74] | Lin X, Zhang Y, Liu H, Boczkaj G, Cao Y, Wang C (2024b) Carbon dioxide sequestration by industrial wastes through mineral carbonation: current status and perspectives. J Clean Prod 434:140258. https://doi.org/10.1016/j.jclepro.2023.140258 |
[75] | Lisabeth H, Zhu W, Xing T, De Andrade V (2017) Dissolution-assisted pattern formation during olivine carbonation. Geophys Res Lett 44(19):9622–9631 |
[76] | Liu W, Teng L, Rohani S, Qin Z, Zhao B, Xu CC, Ren S, Liu Q, Liang B (2021) CO2 mineral carbonation using industrial solid wastes: a review of recent developments. Chem Eng J 416:129093 |
[77] | Loring JS, Thompson CJ, Wang Z, Joly AG, Sklarew DS, Schaef HT, Ilton ES, Rosso KM, Felmy AR (2011) In Situ Infrared Spectroscopic Study of Forsterite Carbonation in Wet Supercritical CO2. Environ Sci Technol 45(14):6204–6210. https://doi.org/10.1021/es201284e |
[78] | Loring JS, Chen J, Bénézeth P, Qafoku O, Ilton ES, Washton NM, Thompson CJ, Martin PF, McGrail BP, Rosso KM, Felmy AR, Schaef HT (2015) Evidence for carbonate surface complexation during forsterite carbonation in wet supercritical carbon dioxide. Langmuir 31(27):7533–7543. https://doi.org/10.1021/acs.langmuir.5b01052 |
[79] | Lu X, Carroll KJ, Turvey CC, Dipple GM (2022) Rate and capacity of cation release from ultramafic mine tailings for carbon capture and storage. Appl Geochem 140:105285. https://doi.org/10.1016/j.apgeochem.2022.105285 |
[80] | Ma J, Li L, Wang H, Du Y, Ma J, Zhang X, Wang Z (2022a) Carbon capture and storage: history and the road ahead. Engineering 14:33–43 |
[81] | Ma M, Guo R, Bing L, Wang J, Yin Y, Zhang W, Niu L, Liu Z, Xi F (2022b) Quantitative analysis of CO2 uptake by alkaline solid wastes in China. J Clean Prod 363:132454. https://doi.org/10.1016/j.jclepro.2022.132454 |
[82] | Marieni C, Voigt M, Clark DE, Gíslason SR, Oelkers EH (2021) Mineralization potential of water-dissolved CO2 and H2S injected into basalts as function of temperature: Freshwater versus Seawater. Int J Greenhouse Gas Control 109:103357. https://doi.org/10.1016/j.ijggc.2021.103357 |
[83] | Matter JM, Kelemen PB (2009) Permanent storage of carbon dioxide in geological reservoirs by mineral carbonation. Nat Geosci 2(12):837–841 |
[84] | Matter JM, Broecker W, Gislason S, Gunnlaugsson E, Oelkers E, Stute M, Sigurdardóttir H, Stefansson A, Alfreðsson H, Aradóttir E (2011) The CarbFix Pilot Project–storing carbon dioxide in basalt. Energy Procedia 4:5579–5585 |
[85] | Matter JM, Stute M, Snæbjörnsdottir SÓ, Oelkers EH, Gislason SR, Aradottir ES, Sigfusson B, Gunnarsson I, Sigurdardottir H, Gunnlaugsson E, Axelsson G, Alfredsson HA, Wolff-Boenisch D, Mesfin K, Taya DF (2016) Rapid carbon mineralization for permanent disposal of anthropogenic carbon dioxide emissions. Science 352(6291):1312–1314. https://doi.org/10.1126/science.aad8132 |
[86] | McGrail BP, Spane FA, Amonette JE, Thompson C, Brown CF (2014a) Injection and monitoring at the Wallula basalt pilot project. Energy Procedia 63:2939–2948 |
[87] | McGrail BP, Schaef HT, Spane FA, Horner JA, Owen AT, Cliff JB, Qafoku O, Thompson CJ, Sullivan EC (2017a) Wallula basalt pilot demonstration project: post-injection results and conclusions. Energy Procedia 114:5783–5790. https://doi.org/10.1016/j.egypro.2017.03.1716 |
[88] | McGrail BP, Schaef HT, Ho AM, Chien YJ, Dooley JJ, Davidson C. (2006) Potential for carbon dioxide sequestration in flood basalts. Journal of Geophysical Research: Solid Earth 111 (B12) |
[89] | McGrail BP, Spane FA, Amonette JE, Thompson CR, Brown CF (2014b) Injection and Monitoring at the Wallula Basalt Pilot Project. In: Paper presented at the 12th international conference on greenhouse gas control technologies, GHGT-12 |
[90] | McGrail BP, Schaef HT, Spane FA, Horner JA, Owen AT, Cliff JB, Qafoku O, Thompson CJ, Sullivan EC (2017b) Wallula basalt pilot demonstration project: post-injection results and conclusions. In: Paper presented at the 13th international conference on greenhouse gas control technologies, GHGT-13 |
[91] | Menzel MD, Urai JL, Ukar E, Hirth G, Schwedt A, Kovács A, Kibkalo L, Kelemen PB (2022) Ductile deformation during carbonation of serpentinized peridotite. Nat Commun 13(1):3478. https://doi.org/10.1038/s41467-022-31049-1 |
[92] | Mergelsberg ST, Rajan BP, Legg BA, Kovarik L, Burton SD, Bowers GM, Bowden ME, Qafoku O, Thompson CJ, Kerisit SN, Ilton ES, Loring JS (2023) Nanoscale Mg-depleted layers slow carbonation of forsterite (Mg2SiO4) when water is limited. Environ Sci Technol Lett 10(1):98–104. https://doi.org/10.1021/acs.estlett.2c00866 |
[93] | Mesfin KG, Wolff-Boenisch D, Gislason SR, Oelkers EH (2023) Effect of cation chloride concentration on the dissolution rates of basaltic glass and labradorite: application to subsurface carbon storage. Minerals 13(5):682. https://doi.org/10.3390/min13050682 |
[94] | Metz B, Davidson O, De Coninck H, Loos M, Meyer L (2005) IPCC special report on carbon dioxide capture and storage. Cambridge University Press, Cambridge |
[95] | Miao E, Du Y, Zheng X, Zhang X, Xiong Z, Zhao Y, Zhang J (2023) CO2 sequestration by direct mineral carbonation of municipal solid waste incinerator fly ash in ammonium salt solution: Performance evaluation and reaction kinetics. Sep Purif Technol 309:123103. https://doi.org/10.1016/j.seppur.2023.123103 |
[96] | Miller QRS, Thompson CJ, Loring JS, Windisch CF, Bowden ME, Hoyt DW, Hu JZ, Arey BW, Rosso KM, Schaef HT (2013) Insights into silicate carbonation processes in water-bearing supercritical CO2 fluids. Int J Greenhouse Gas Control 15:104–118. https://doi.org/10.1016/j.ijggc.2013.02.005 |
[97] | Miller QRS, Kaszuba JP, Schaef HT, Bowden ME, McGrail BP (2015) Impacts of organic ligands on forsterite reactivity in supercritical CO2 fluids. Environ Sci Technol 49(7):4724–4734. https://doi.org/10.1021/es506065d |
[98] | Miller QR, Schaef HT, Kaszuba JP, Qiu L, Bowden ME, McGrail BP (2018a) Tunable manipulation of mineral carbonation kinetics in nanoscale water films via citrate additives. Environ Sci Technol 52(12):7138–7148 |
[99] | Miller QRS, Ilton ES, Qafoku O, Dixon DA, Vasiliu M, Thompson CJ, Schaef HT, Rosso KM, Loring JS (2018b) Water structure controls carbonic acid formation in adsorbed water films. J Phys Chem Lett 9(17):4988–4994. https://doi.org/10.1021/acs.jpclett.8b02162 |
[100] | Miller QRS, Schaef HT, Kaszuba JP, Qiu L, Bowden ME, McGrail BP (2018c) Tunable manipulation of mineral carbonation kinetics in nanoscale water films via citrate additives. Environ Sci Technol 52(12):7138–7148. https://doi.org/10.1021/acs.est.8b00438 |
[101] | Miller QR, Kaszuba JP, Schaef HT, Bowden ME, McGrail BP, Rosso KM (2019a) Anomalously low activation energy of nanoconfined MgCO3 precipitation. Chem Commun 55(48):6835–6837 |
[102] | Miller QRS, Dixon DA, Burton SD, Walter ED, Hoyt DW, McNeill AS, Moon JD, Thanthiriwatte KS, Ilton ES, Qafoku O, Thompson CJ, Schaef HT, Rosso KM, Loring JS (2019b) Surface-catalyzed oxygen exchange during mineral carbonation in nanoscale water films. J Phys Chem C 123(20):12871–12885. https://doi.org/10.1021/acs.jpcc.9b02215 |
[103] | Miller QRS, Schaef HT, Kaszuba JP, Gadikota G, McGrail BP, Rosso KM (2019c) Quantitative review of olivine carbonation kinetics: reactivity trends, mechanistic insights, and research frontiers. Environ Sci Technol Lett 6(8):431–442. https://doi.org/10.1021/acs.estlett.9b00301 |
[104] | Miller QRS, Kaszuba JP, Kerisit SN, Schaef HT, Bowden ME, McGrail BP, Rosso KM (2020) Emerging investigator series: ion diffusivities in nanoconfined interfacial water films contribute to mineral carbonation thresholds. Environ Sci Nano 7(4):1068–1081. https://doi.org/10.1039/C9EN01382B |
[105] | Mishra A, Boon MM, Benson SM, Watson MN, Haese RR (2023) Reconciling predicted and observed carbon mineralization in siliciclastic formations. Chem Geol 619:121324. https://doi.org/10.1016/j.chemgeo.2023.121324 |
[106] | Mito S, Xue Z, Ohsumi T (2008) Case study of geochemical reactions at the Nagaoka CO2 injection site, Japan. Int J Greenhouse Gas Control 2(3):309–318 |
[107] | NASEM (2019) Negative emissions technologies and reliable sequestration: a research agenda. National Academies Press |
[108] | Noiriel C, Daval D (2017) Pore-scale geochemical reactivity associated with CO2 storage: new frontiers at the fluid-solid interface. Acc Chem Res 50(4):759–768. https://doi.org/10.1021/acs.accounts.7b00019 |
[109] | Oelkers EH (2001) General kinetic description of multioxide silicate mineral and glass dissolution. Geochim Cosmochim Acta 65(21):3703–3719 |
[110] | Oelkers EH, Gislason SR, Matter J (2008) Mineral carbonation of CO2. Elements 4(5):333–337. https://doi.org/10.2113/gselements.4.5.333 |
[111] | Oelkers EH, Declercq J, Saldi GD, Gislason SR, Schott J (2018) Olivine dissolution rates: a critical review. Chem Geol 500:1–19 |
[112] | Oelkers EH, Arkadakskiy S, Afifi AM, Hoteit H, Richards M, Fedorik J, Delaunay A, Torres JE, Ahmed ZT, Kunnummal N (2022) The subsurface carbonation potential of basaltic rocks from the Jizan region of Southwest Saudi Arabia. Int J Greenhouse Gas Control 120:103772 |
[113] | Orr FM Jr (2009) Onshore geologic storage of CO2. Science 325(5948):1656–1658 |
[114] | Oskierski HC, Turvey CC, Wilson SA, Dlugogorski BZ, Altarawneh M, Mavromatis V (2021) Mineralisation of atmospheric CO2 in hydromagnesite in ultramafic mine tailings - Insights from Mg isotopes. Geochim Cosmochim Acta 309:191–208. https://doi.org/10.1016/j.gca.2021.06.020 |
[115] | Osselin F, Pichavant M, Champallier R, Ulrich M, Raimbourg H (2022) Reactive transport experiments of coupled carbonation and serpentinization in a natural serpentinite. Implication for hydrogen production and carbon geological storage. Geochim Cosmochim Acta 318:165–189. https://doi.org/10.1016/j.gca.2021.11.039 |
[116] | Ostovari H, Müller L, Mayer F, Bardow A (2022) A climate-optimal supply chain for CO2 capture, utilization, and storage by mineralization. J Clean Prod 360:131750. https://doi.org/10.1016/j.jclepro.2022.131750 |
[117] | Peuble S, Godard M, Luquot L, Andreani M, Martinez I, Gouze P (2015) CO2 geological storage in olivine rich basaltic aquifers: New insights from reactive-percolation experiments. Appl Geochem 52:174–190. https://doi.org/10.1016/j.apgeochem.2014.11.024 |
[118] | Peuble S, Andreani M, Gouze P, Pollet-Villard M, Reynard B, Van de Moortele B (2018) Multi-scale characterization of the incipient carbonation of peridotite. Chem Geol 476:150–160. https://doi.org/10.1016/j.chemgeo.2017.11.013 |
[119] | Peuble S, Godard M, Gouze P, Leprovost R, Martinez I, Shilobreeva S (2019) Control of CO2 on flow and reaction paths in olivine-dominated basements: an experimental study. Geochim Cosmochim Acta 252:16–38. https://doi.org/10.1016/j.gca.2019.02.007 |
[120] | Placencia-Gómez E, Kerisit SN, Mehta HS, Qafoku O, Thompson CJ, Graham TR, Ilton ES, Loring JS (2020) Critical water coverage during forsterite carbonation in thin water films: activating dissolution and mass transport. Environ Sci Technol 54(11):6888–6899. https://doi.org/10.1021/acs.est.0c00897 |
[121] | Pogge von Strandmann PAE, Burton KW, Snæbjörnsdóttir SO, Sigfússon B, Aradóttir ES, Gunnarsson I, Alfredsson HA, Mesfin KG, Oelkers EH, Gislason SR (2019) Rapid CO2 mineralisation into calcite at the CarbFix storage site quantified using calcium isotopes. Nat Commun 10(1):1983. https://doi.org/10.1038/s41467-019-10003-8 |
[122] | Polites EG, Schaef HT, Horner JA, Owen AT, Holliman JE Jr, McGrail BP, Miller QRS (2022) Exotic carbonate mineralization recovered from a deep basalt carbon storage demonstration. Environ Sci Technol 56(20):14713–14722. https://doi.org/10.1021/acs.est.2c03269 |
[123] | Power IM, Harrison AL, Dipple GM, Wilson S, Kelemen PB, Hitch M, Southam G (2013) Carbon mineralization: from natural analogues to engineered systems. Rev Mineral Geochem 77(1):305–360 |
[124] | Qin Z, He Y, Wang L, Liu X, Ding Y, Chen Z, Li B, Wang Z (2024) Analytical model for assessing the impact of high-rate CO2 injection on dry-out and salt precipitation in the near-wellbore region of saline aquifers. Energy Fuels 38(10):8875–8894. https://doi.org/10.1021/acs.energyfuels.4c00153 |
[125] | Ramos GMS, Barbosa JA, de Araújo AFL, Correia Filho OJ, Barreto CJS, Oliveira JTC, de Medeiros RS (2023) Potential for permanent CO2 sequestration in depleted volcanic reservoirs in the offshore Campos basin, Brazil. Int J Greenhouse Gas Control 128:103942 |
[126] | Rashid MI, Benhelal E, Anderberg L, Farhang F, Oliver T, Rayson MS, Stockenhuber M (2022) Aqueous carbonation of peridotites for carbon utilisation: a critical review. Environ Sci Pollut Res 29(50):75161–75183. https://doi.org/10.1007/s11356-022-23116-3 |
[127] | Rasool MH, Ahmad M (2023) Reactivity of basaltic minerals for CO2 sequestration via in situ mineralization: a review. Minerals 13(9):1154 |
[128] | Raza A, Glatz G, Gholami R, Mahmoud M, Alafnan S (2022) Carbon mineralization and geological storage of CO2 in basalt: mechanisms and technical challenges. Earth Sci Rev 229:104036 |
[129] | Reynes JF, Mercier G, Blais J-F, Pasquier L-C (2023) Carbon dioxide sequestration by mineral carbonation via iron complexation using bipyridine chelating ligands. Dalton Trans 52(19):6536–6542. https://doi.org/10.1039/D3DT00563A |
[130] | Rimstidt JD, Brantley SL, Olsen AA (2012) Systematic review of forsterite dissolution rate data. Geochim Cosmochim Acta 99:159–178 |
[131] | Rosenqvist MP, Meakins MWJ, Planke S, Millett JM, Kjøll HJ, Voigt MJ, Jamtveit B (2023) Reservoir properties and reactivity of the faroe islands basalt group: investigating the potential for CO2 storage in the North Atlantic Igneous Province. Int J Greenhouse Gas Control 123:103838. https://doi.org/10.1016/j.ijggc.2023.103838 |
[132] | Ruiz-Lopez MF, Francisco JS, Martins-Costa MT, Anglada JM (2020) Molecular reactions at aqueous interfaces. Nat Rev Chem 4(9):459–475 |
[133] | Saldi GD, Daval D, Morvan G, Knauss KG (2013) The role of Fe and redox conditions in olivine carbonation rates: An experimental study of the rate limiting reactions at 90 and 150°C in open and closed systems. Geochim Cosmochim Acta 118:157–183. https://doi.org/10.1016/j.gca.2013.04.029 |
[134] | Saldi GD, Daval D, Guo H, Guyot F, Bernard S, Guillou CL, Davis JA, Knauss KG (2015) Mineralogical evolution of Fe–Si-rich layers at the olivine-water interface during carbonation reactions. Am Miner 100(11–12):2655–2669. https://doi.org/10.2138/am-2015-5340 |
[135] | Sandalow, D., Aines, R., Friedmann, J., Kelemen, P., McCormick, C., Power, I., Schmidt, B., Wilson, S. (2021) 'Carbon mineralization roadmap draft october 2021'. Lawrence Livermore National Lab. (LLNL), Livermore, CA (United States). |
[136] | Sanna A, Maroto-valer M (2016) CO2 sequestration by ex-situ mineral carbonation. World Scientific Publishing Company |
[137] | Sanna A, Uibu M, Caramanna G, Kuusik R, Maroto-Valer M (2014) A review of mineral carbonation technologies to sequester CO2. Chem Soc Rev 43(23):8049–8080 |
[138] | Santos HS, Nguyen H, Venâncio F, Ramteke D, Zevenhoven R, Kinnunen P (2023) Mechanisms of Mg carbonates precipitation and implications for CO2 capture and utilization/storage. Inorg Chem Front 10(9):2507–2546 |
[139] | Schaef HT, Miller QRS, Thompson CJ, Loring JS, Bowden MS, Arey BW, McGrail BP, Rosso KM (2013) Silicate carbonation in supercritical CO2 containing dissolved H2O: an in situ high pressure X-ray diffraction and infrared spectroscopy study. Energy Procedia 37:5892–5896. https://doi.org/10.1016/j.egypro.2013.06.514 |
[140] | Scheifele B, Saika-Voivod I, Bowles RK, Poole PH (2013) Heterogeneous nucleation in the low-barrier regime. Phys Rev E 87(4):042407 |
[141] | Schott J, Pokrovsky OS, Oelkers EH (2009) The link between mineral dissolution/precipitation kinetics and solution chemistry. Rev Mineral Geochem 70(1):207–258 |
[142] | Seifritz W (1990) CO2 disposal by means of silicates. Nature 345:486–486 |
[143] | Sigfusson B, Gislason SR, Matter JM, Stute M, Gunnlaugsson E, Gunnarsson I, Aradottir ES, Sigurdardottir H, Mesfin K, Alfredsson HA (2015) Solving the carbon-dioxide buoyancy challenge: The design and field testing of a dissolved CO2 injection system. Int J Greenhouse Gas Control 37:213–219 |
[144] | Sissmann O, Daval D, Brunet F, Guyot F, Verlaguet A, Pinquier Y, Findling N, Martinez I (2013) The deleterious effect of secondary phases on olivine carbonation yield: Insight from time-resolved aqueous-fluid sampling and FIB-TEM characterization. Chem Geol 357:186–202. https://doi.org/10.1016/j.chemgeo.2013.08.031 |
[145] | Snæbjörnsdóttir SÓ, Gislason SR (2016) CO2 storage potential of basaltic rocks offshore Iceland. Energy Procedia 86:371–380 |
[146] | Snæbjörnsdóttir SÓ, Wiese F, Fridriksson T, Ármansson H, Einarsson GM, Gislason SR (2014) CO2 storage potential of basaltic rocks in Iceland and the oceanic ridges. Energy Procedia 63:4585–4600 |
[147] | Snæbjörnsdóttir SÓ, Oelkers EH, Mesfin K, Aradóttir ES, Dideriksen K, Gunnarsson I, Gunnlaugsson E, Matter JM, Stute M, Gislason SR (2017) The chemistry and saturation states of subsurface fluids during the in situ mineralisation of CO2 and H2S at the CarbFix site in SW-Iceland. Int J Greenhouse Gas Control 58:87–102 |
[148] | Snæbjörnsdóttir SÓ, Sigfússon B, Marieni C, Goldberg D, Gislason SR, Oelkers EH (2020) Carbon dioxide storage through mineral carbonation. Nat Rev Earth Environ 1(2):90–102 |
[149] | Soulaine C, Roman S, Kovscek A, Tchelepi HA (2017) Mineral dissolution and wormholing from a pore-scale perspective. J Fluid Mech 827:457–483 |
[150] | Stillings M, Shipton ZK, Lunn RJ (2023) Mechanochemical processing of silicate rocks to trap CO2. Nat Sustain 6(7):780–788. https://doi.org/10.1038/s41893-023-01083-y |
[151] | Sun J, Bertos MF, Simons SJ (2008) Kinetic study of accelerated carbonation of municipal solid waste incinerator air pollution control residues for sequestration of flue gas CO2. Energy Environ Sci 1(3):370–377 |
[152] | Sun L, Liu Y, Cheng Z, Jiang L, Lv P, Song Y (2023) Review on Multiscale CO2 mineralization and geological storage: mechanisms, characterization, modeling. Appl Persp Energy Fuels 37(19):14512–14537 |
[153] | Tencent (2023) Remove Carbon Permanently with Carbonx. https://carbonx.world/ |
[154] | Thompson CJ, Loring JS, Rosso KM, Wang Z (2013) Comparative reactivity study of forsterite and antigorite in wet supercritical CO2 by in situ infrared spectroscopy. Int J Greenhouse Gas Control 18:246–255. https://doi.org/10.1016/j.ijggc.2013.07.007 |
[155] | Todd Schaef H, McGrail BP, Loring JL, Bowden ME, Arey BW, Rosso KM (2013) Forsterite [Mg2SiO4)] carbonation in wet supercritical CO2: an in situ high-pressure X-ray diffraction study. Environ Sci Technol 47(1):174–181. https://doi.org/10.1021/es301126f |
[156] | Tokunaga TK, Finsterle S, Kim Y, Wan J, Lanzirotti A, Newville M (2017) Ion diffusion within water films in unsaturated porous media. Environ Sci Technol 51(8):4338–4346 |
[157] | Turvey CC, Wilson S, Hamilton JL, Tait AW, McCutcheon J, Beinlich A, Fallon SJ, Dipple GM, Southam G (2018) Hydrotalcites and hydrated Mg-carbonates as carbon sinks in serpentinite mineral wastes from the Woodsreef chrysotile mine, New South Wales, Australia: controls on carbonate mineralogy and efficiency of CO2 air capture in mine tailings. Int J Greenhouse Gas Control 79:38–60 |
[158] | Tutolo BM, Awolayo A, Brown C (2021) Alkalinity generation constraints on basalt carbonation for carbon dioxide removal at the Gigaton-per-year scale. Environ Sci Technol 55(17):11906–11915. https://doi.org/10.1021/acs.est.1c02733 |
[159] | UNEP (2022) 'Emissions Gap Report 2022: The Closing Window — Climate crisis calls for rapid transformation of societies'. Nairobi |
[160] | Van Noort R, Spiers C, Drury M, Kandianis M (2013) Peridotite dissolution and carbonation rates at fracture surfaces under conditions relevant for in situ mineralization of CO2. Geochim Cosmochim Acta 106:1–24 |
[161] | Wallace AF, Hedges LO, Fernandez-Martinez A, Raiteri P, Gale JD, Waychunas GA, Whitelam S, Banfield JF, De Yoreo JJ (2013) Microscopic evidence for liquid-liquid separation in supersaturated CaCO3 solutions. Science 341(6148):885–889 |
[162] | Wang T, Huang H, Hu X, Fang M, Luo Z, Guo R (2017) Accelerated mineral carbonation curing of cement paste for CO2 sequestration and enhanced properties of blended calcium silicate. Chem Eng J 323:320–329 |
[163] | Wang F, Dreisinger D, Jarvis M, Hitchins T (2021a) Kinetic evaluation of mineral carbonation of natural silicate samples. Chem Eng J 404:126522 |
[164] | Wang T, Yi Z, Guo R, Huang H, Garcia S, Maroto-Valer MM (2021b) Particle carbonation kinetics models and activation methods under mild environment: the case of calcium silicate. Chem Eng J 423:130157. https://doi.org/10.1016/j.cej.2021.130157 |
[165] | Wang Y, Liu JH, Hu X, Chang J, Zhang TT, Shi CJ (2023) Utilization of accelerated carbonation to enhance the application of steel slag: a review. J Sustain Cement-Based Mater 12(4):471–486. https://doi.org/10.1080/21650373.2022.2154287 |
[166] | Wei Y-M, Kang J-N, Liu L-C, Li Q, Wang P-T, Hou J-J, Liang Q-M, Liao H, Huang S-F, Yu B (2021) A proposed global layout of carbon capture and storage in line with a 2 °C climate target. Nat Clim Chang 11(2):112–118. https://doi.org/10.1038/s41558-020-00960-0 |
[167] | White SK, Spane FA, Schaef HT, Miller QRS, White MD, Horner JA, McGrail BP (2020) Quantification of CO2 mineralization at the Wallula basalt pilot project. Environ Sci Technol 54(22):14609–14616. https://doi.org/10.1021/acs.est.0c05142 |
[168] | White S, Xue Z, Sato T, Hong Y (2006) Modeling and analysis of the pressure response in the CO2 injection experiment conducted at Nagaoka, Japan. Proc. GHGT-8 1 242–243 |
[169] | Wiese F, Fridriksson T, Ármannsson H (2008) CO2 fixation by calcite in high-temperature geothermal systems in Iceland. Report from the Iceland Geosurvey (Ísor), Ísor-2008/003, Reykjavik |
[170] | Wood CE, Qafoku O, Loring JS, Chaka AM (2019) Role of Fe (II) content in olivine carbonation in wet supercritical CO2. Environ Sci Technol Lett 6(10):592–599 |
[171] | Wu X, Qu A, Gao P, Pang W, Li Q, Zhao J, Yang L, Song Y (2024) The Future of CO2 geological sequestration: an overview of mineralization in aqueous and supercritical states. Energy Fuels 38(11):9320–9338. https://doi.org/10.1021/acs.energyfuels.4c01139 |
[172] | Xu R, Li R, Ma J, He D, Jiang P (2017) Effect of mineral dissolution/precipitation and CO2 exsolution on CO2 transport in geological carbon storage. Acc Chem Res 50(9):2056–2066 |
[173] | Yadav S, Mehra A (2019) Mathematical modelling and experimental study of carbonation of wollastonite in the aqueous media. J CO2 Util 31:181–191 |
[174] | Yadav S, Mehra A (2021) A review on ex situ mineral carbonation. Environ Sci Pollut Res 28:12202–12231 |
[175] | Yamamoto H, Nakajima T, Xue Z (2017) Quantitative interpretation of trapping mechanisms of CO2 at Nagaoka pilot project a history matching study for 10-year post-injection–. Energy Procedia 114:5058–5069 |
[176] | Yang L, Xu T, Yang B, Tian H, Lei H (2014) Effects of mineral composition and heterogeneity on the reservoir quality evolution with CO2 intrusion. Geochem Geophys Geosyst 15(3):605–618. https://doi.org/10.1002/2013GC005157 |
[177] | Zakharova NV, Goldberg DS, Sullivan EC, Herron MM, Grau JA (2012) Petrophysical and geochemical properties of Columbia River flood basalt: implications for carbon sequestration. Geochem Geophys Geosyst. https://doi.org/10.1029/2012GC004305 |
[178] | Zare S, Funk A, Abdolhosseini Qomi MJ (2022a) Formation and dissolution of surface metal carbonate complexes: implications for interfacial carbon mineralization in metal silicates. J Phys Chem C 126(28):11574–11584. https://doi.org/10.1021/acs.jpcc.2c02981 |
[179] | Zare S, Uddin KMS, Funk A, Miller QRS, Abdolhosseini Qomi MJ (2022b) Nanoconfinement matters in humidified CO2 interaction with metal silicates. Environ Sci Nano 9(10):3766–3779. https://doi.org/10.1039/D2EN00148A |
[180] | Zhang S, DePaolo DJ (2017) Rates of CO2 mineralization in geological carbon storage. Acc Chem Res 50(9):2075–2084 |
[181] | Zhang Q, Tutolo BM (2022) Evaluation of the potential of glauconite in the Western Canadian Sedimentary Basin for large-scale carbon dioxide mineralization. Int J Greenhouse Gas Control 117:103663. https://doi.org/10.1016/j.ijggc.2022.103663 |
[182] | Zhang SY, Li H, Cai J, Olabi AG, Anthony EJ, Manovic V (2020) Recent advances in carbon dioxide utilization. Renew Sustain Energy Rev 125:109799. https://doi.org/10.1016/j.rser.2020.109799 |
[183] | Zhang L, Chen L, Hu R, Cai J (2022) Subsurface multiphase reactive flow in geologic CO2 storage: Key impact factors and characterization approaches. Adv Geo-Energy Res 6(3):179–180 |
[184] | Zhang L, Wen R, Li F, Li C, Sun Y, Yang H (2023) Assessment of CO2 mineral storage potential in the terrestrial basalts of China. Fuel 348:128602. https://doi.org/10.1016/j.fuel.2023.128602 |
13 April 2024
01 July 2024
09 January 2025
November -0001
https://doi.org/10.1007/s40789-025-00755-8