Next Article in Journal
Association Between TP53 Mutations and Platinum Resistance in a Cohort of High-Grade Serous Ovarian Cancer Patients: Novel Implications for Personalized Therapeutics
Previous Article in Journal
Molecular Biomarkers in Neurological Diseases: Advances in Diagnosis and Prognosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mineral Carbonation for Carbon Sequestration: A Case for MCP and MICP

by
Samantha M. Wilcox
1,
Catherine N. Mulligan
1,* and
Carmen Mihaela Neculita
2
1
Department of Building, Civil and Environmental Engineering, Concordia University, Montréal, QC H3G IM8, Canada
2
Research Institute on Mines and the Environment (RIME), University of Quebec in Abitibi-Témiscamingue, Rouyn-Noranda, QC J9X 5E4, Canada
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(5), 2230; https://doi.org/10.3390/ijms26052230
Submission received: 31 January 2025 / Revised: 24 February 2025 / Accepted: 27 February 2025 / Published: 1 March 2025
(This article belongs to the Section Biochemistry)

Abstract

:
Mineral carbonation is a prominent method for carbon sequestration. Atmospheric carbon dioxide (CO2) is trapped as mineral carbonate precipitates, which are geochemically, geologically, and thermodynamically stable. Carbonate rocks can originate from biogenic or abiogenic origin, whereby the former refers to the breakdown of biofragments and the latter precipitation out of water. Carbonates can also be formed through biologically controlled mechanisms (BCMs), biologically mediated mechanisms (BMMs), and biologically induced mechanisms (BIMs). Microbial carbonate precipitation (MCP) is a BMM occurring through the interaction of organics (extracellular polymeric substances (EPS), cell wall, etc.) and soluble cations facilitating indirect precipitation of carbonate minerals. Microbially induced carbonate precipitation (MICP) is a BIM occurring via different metabolic pathways. Enzyme-driven pathways (carbonic anhydrase (CA) and/or urease), specifically, are promising for the high conversion to calcium carbonate (CaCO3) precipitation, trapping large quantities of gaseous CO2. These carbonate precipitates can trap CO2 via mineral trapping, solubility trapping, and formation trapping and aid in CO2 leakage reduction in geologic carbon sequestration. Additional experimental research is required to assess the feasibility of MICP for carbon sequestration at large scale for long-term stability of precipitates. Laboratory-scale evaluation can provide preliminary data on preferable metabolic pathways for different materials and their capacity for carbonate precipitation via atmospheric CO2 versus injected CO2.

1. Introduction

Global warming and climate change have been significant concerns to scientists, engineers, and policy makers for a long time. The first Intergovernmental Panel on Climate Change (IPCC) was held in 1988 and the notorious Paris Agreement in 2015, which involved a commitment by 195 countries to limit global warming to 1.5–2 °C. Today, despite efforts to reduce greenhouse gas (GHG) emissions, global warming is projected at 3.5 °C by 2100 [1]. There is a need for carbon sequestration strategies that transform and sequester GHGs (CO2, methane (CH4) and fluorinated gases (hydrofluorocarbons (HFCs), perfluorinated compounds (PFCs), sulfur hexafluoride (SF6), and nitrogen trifluoride (NF3)) from the atmosphere, which, left untreated, create a warming effect.
Biological carbon sequestration can utilize plants and microorganisms (bacteria, fungi, archaea, cyanobacteria, and algae) to fix inorganic CO2 as organic products (cellulous, lignocellulose, chitin, hemicellulose, lignin, etc.) or mineral precipitates for carbon capture and utilization [2]. Some microorganisms (carboxydotrophs) are carbon dependent and utilize atmospheric CO and CO2 as their energy source [3]. Further, there are many identified pathways and enzymes (e.g., CA, RuBisCO, carbon monoxide dehydrogenase (CODH), etc.) linked to biologic carbon sequestration [2]. Mineral carbonation, specifically, biogenic mineral carbonation, offers a promising opportunity to sequester atmospheric CO2 through naturally occurring processes. The process can be applied passively (trapping atmospheric CO2) or as a carbon capture and storage (CCS; actively injecting CO2) strategy.
While microbially induced carbonate precipitation (MICP) is a well-studied biological technique for soil and cement strengthening and restoration of construction materials, limited research has evaluated its feasibility as a carbon sequestration technique. In short, the microorganisms act as a catalyst to chemical precipitation and capture carbon as mineral carbonate precipitates. These reactions can be naturally occurring or engineered to enhance or optimize precipitation and therefore carbon sequestration. Alternatively, microbial carbonate precipitation (MCP) is a passive precipitation technique driven by the organic material in the environment. The objective of this paper is to evaluate abiotic and biotic carbonation, making a case for MICP and MCP as viable mineral carbonation techniques for carbon sequestration.

2. Biochemical Precipitation

2.1. Carbonate Precipitates

Chemical precipitation is a complex process used for separation of solid substances from solution [4,5,6]. For precipitation to occur, the solute concentration must exceed the liquid–solid equilibrium of the solution, meaning it is in a supersaturated state [5]. Supersaturation is, therefore, the driving force of precipitation [4,5]. Thermodynamically, the Gibbs free energy (G) and the solubility product constant (Ksp) govern the reaction equilibria and the solubility equilibria, respectively [7].
Kinetically, there are three main processes that direct precipitation. These include nucleation, growth, and agglomeration [5]. Nucleation refers to the birth of particles via condensation of ions. It can be homogenous (spontaneous) or heterogenous (prompted by foreign particles) [5,7]. Growth and agglomeration refer to the enlargement of particles. Crystal growth refers to the enlargement from material deposition onto formed particles via transport to the crystal surface, adsorption onto the crystal surface, or formation of crystal lattice bonds [7]. Agglomeration refers to the contact of two or more particles, which over time forms a stable particle [5]. The supersaturation will impact the type of nucleation and growth, which affects the texture and purity of the crystals [5]. Precipitates can also undergo aging (re-arrangement of the crystal structure to form larger, pure crystals with time) and coprecipitation (ion inclusion into the crystal structure or adsorption onto the crystal surface) [7]. Chemical precipitation factors include the soil–water system, pH (favors high pH), Eh, type and concentration of metals and metalloids (metal(loid)s), dissolved organic carbon (DOC), and inorganic and organic ligand presence [8].
Carbonate rocks can exist as igneous [9], sedimentary [10], and metamorphic rocks [11]. However, they primarily present as sedimentary rocks, either as limestone (CaCO3; calcite and aragonite) or dolostone (CaMg(CO3)2) [12]. However, minerals are distinct from rocks, and there are numerous carbonate minerals: magnesite (MgCO3), siderite (FeCO3), dolomite (CaMg(CO3)2), otavite (CdCO3), rhodochrosite (MnCO3), cerussite (PbCO3), smithsonite (ZnCO3), strontianite (SrCO3), witherite (BaCO3), etc. Calcium carbonate is the most abundant of the carbonate minerals, occurring as calcite, aragonite, and vaterite anhydrous polymorphs [10,12], whereby calcite is the most thermodynamically stable and vaterite is the least. Less reported are the hydrous (monoclinic ikaite (CaCO3 · 6H2O) and calcium carbonate monohydrate (CaCO3 · H2O)) and amorphous (ACC–CaCO3 · nH2O) polymorphs [12].
The crystal structure and morphology of carbonate minerals varies significantly. Calcite and dolomite typically have a hexagonal crystallography, whereas aragonite is usually an orthorhombic structure [13,14]. However, the degree of supersaturation has been shown to alter the crystal form [15]. This is in conjunction with morphologic changes in the minerals due to crystal growth rate dependent on supersaturation [12,16]. Therefore, the solubility of carbonate minerals is essential to crystal formation. The reversible reaction for CaCO3 dissolution and precipitation is shown in Equation (1), where the rightward reaction illustrates the dissolution of carbonate minerals, and the leftward reaction demonstrates precipitation [12]. Relative solubility is impacted by crystal size, heterogeneity, defects, porosity, and organic matrices, with mineralogy the most significant [12]. Magnesium (Mg2+), for example, is known to inhibit calcite growth [16,17,18] and increase calcite solubility [12]. Furthermore, CO2 concentration impacts solubility and precipitation, in which an increase in CO2 also increases CaCO3 solubility, causing dissolution [12,19]. Temperature and pressure also influence calcite formation, since temperature and pressure increases can cause CaCO3 solubility decreases, which favors precipitation [12,19,20]. It should be noted that CO2 solubility is inversely correlated to temperature, and positively correlated to pressure [21].
C a C O 3 + H 2 O + C O 2 C a 2 + + 2 H C O 3
Carbonate minerals form carbonate rocks through deposition and diagenesis [12]. They are formed in marine (ocean, sea (neritic and pelagic)) and terrestrial (lakes, hot and cold springs, caves, soils) environments, originating from biogenic, abiogenic, or complex mixtures of both components [12]. Carbonate sediments can be categorized as deep-sea oozes, carbonate turbidites, shelf accumulations of lime sands, silts, muds, organic reefs, and reef debris [22]. They are classified according to their compositions, fabric, and origin [23]. The biosphere and depositional environment impact the skeleton mineralogic, petrographic, and geochemical vestiges of the carbonate rocks [12]. Biogenic deposition often refers to the breakdown of invertebrate biofragments (i.e., shells, single cells, colonial skeletons) and crystallites within algal tissue or the calcification of microbes (i.e., calcimicrobes), whereas abiogenic deposition results from precipitation out of seawater or freshwater [12]. Diagenesis is a complex process encompassing 30 processes, including lithification destructive processes, re-crystallization, and grain-diminution [24]. In Precambrian times, carbonate rocks originated from algae pH control in lagoons and direct chemical precipitation out of sea water, while Cambrian origin often resulted from the organisms extracting carbonate out of sea water [25].

2.2. Biogenic Carbonate Precipitates

Bioprecipitation can encompass the formation of all biologically facilitated crystalline or amorphous precipitates with both organic and/or inorganic components. Bioprecipitation uses microorganisms to catalyze chemical precipitation reactions. It can incorporate different microorganisms to facilitate distinct metabolic pathways. The aim is to precipitate compounds (i.e., carbonates, hydroxides, phosphates, sulfides, sulfates, arsenates, silicas, chlorides, fluorides, oxides, oxalates, etc. [26,27,28,29]) with low solubility.
Abiotic precipitation varies significantly from biotic precipitation. The morphology is a key indicator used to distinguish inorganic and abiotic processes from biogenic minerals [30], which are typically differentiated by their unusual external morphology [29]. An interesting characteristic of biominerals is the composites or agglomeration of crystals separated by organic material [29]. Researchers have found differences in shape, size, crystallinity, isotopic, trace compositions, organic functional groups, activation energy, and enthalpy between biotic and abiotic precipitates [29,30]. Within lacustrine systems, for example, carbonate minerals are typically of biological origin or from direct biological activity, whereby the oxic (bio-induced pelagic CO32− precipitates), suboxic, and anoxic (microbial-induced diagenetic CO32− precipitates) conditions in the microenvironment with the ion supply impact the microbial pathway (see Section 2.4) [31]. Therefore, microenvironmental conditions impact the carbon isotopic composition of dissolved inorganic carbon (DIC) in pore water and the carbon isotopic composition of precipitates [31]. Further differences have been reported between biotic, organogenic (nutrient composition without bacterial cells), and inorganogenic (chemical reaction) forms of CaCO3 precipitation, suggesting thermal stability is highest in biotic calcite [30]. The calcite crystal growth rate and biotic growth rate in carbonate deposits will influence whether it is biotic/abiotic [32]. If the supersaturation state is high, calcite will favor abiotic precipitation, as crystal formation outpaces microbial growth rates [32]. In the context of carbon sequestration, abiotic factors regulating sequestration include pH and medium components (i.e., urea), while biotic factors are dependent on the species or strains [33]. Again, resident biota will impact the CO2 levels inducing dissolution (water absorption of CO2 respiration in soil increasing acidity and dissolution) or precipitation (removal of CO2 in seawater via phototrophs during photosynthesis) [12]. The degree of control exerted by the microorganism will dictate the biological mechanism occurring [29].
The interaction between microbial activity, the external environmental conditions, and the overall biofilm matrix will determine how and whether biotic precipitates form. There are three primary mechanisms (Figure 1 and Table 1) capable of facilitating bioprecipitation, including biologically controlled mechanisms (BCMs), biologically induced mechanisms (BIMs), and biologically mediated mechanisms (BMMs; otherwise termed biologically influenced mechanisms). Under certain conditions, microorganisms can synthesize minerals via nucleation and growth facilitating BCM [26,27,29,34,35]. Cellular activities including active pumping, passive diffusion, and secretion can lead to precipitation of particles in the extracellular, intracellular, or intercellular environment [26,27,29]. The final resting place of precipitates is within or on the microbial cell [27,34,36]. The composition, morphology, and localization of precipitates are influenced by the species-specific process [26,29]. BIMs involve the metabolic activity of the microorganism, which interact with the environment to facilitate precipitation [26,27,29,36]. The precipitates form in the extracellular environment, where nucleation and growth typically transpire on the microbial cell wall [29]. Precipitation is dependent on the environmental conditions (i.e., pH, redox potential, CO2, etc.) and the subsequent supersaturation [29,34,35,37]. The composition, particle size, crystal purity, and morphology are varied due to diverse environmental conditions [26,29]. Passive precipitation is caused by the BMM due to the interaction of organic matter (i.e., extracellular polymeric substances (EPS), biofilm, and the organic/inorganic compounds) within the matrix [26,34]. Biological activity does not directly cause precipitation.
Both BIM and BMM utilize prokaryotes to facilitate precipitation. While the presence of microorganisms is not directly required for BMM, the organic EPS matrix is an extension of the microbial cell [38], as shown in Figure 1. Mineral deposits via BIM and BMM can be classified as stromatolites, thrombolites, and leiolite [38]. For carbonate precipitation, the BIM and BMM are referred to as MICP and MCP, respectively. Table 2 provides a comparative analysis of these methods. For global carbon sequestration, MCP is well established as a long-term storage technique for carbon. Carbon can be trapped in terrestrial environments (i.e., soils, caves, deserts, tundra, boreal forests, temperate forests, tropical forests, grasslands, etc.) in the organic matrix of soil as soil organic carbon (SOC), carbonate deposits (precipitated via plants, fungi, or bacteria), and/or vegetation [38,39,40]. Plants are also able to store inorganic CO2 through biological carbon mitigation as organic carbon through photosynthesis [41]. As a byproduct of photosynthesis, plants can precipitate whewellite (Ca(C2O4) · H2O) storing atmospheric CO2 [38]. Liu et al. [42] found that EPS and EPS-carbon are positively correlated to SOC, whereby EPS-carbon accounts for ≤10.69% the total organic carbon in surface sediments. This is quite significant, since 75% of organic carbon is sequestered in mangrove ecosystems in sediments [42,43]. MICP, however, is less established in terms of global carbon sequestration. MICP is shown to store carbon as carbonate deposits in marine environments, hypersaline lakes, freshwater environments, and continental environments [38], which act as carbon sinks for carbon sequestration.
Table 1. Comparative analysis of BCM, BIM, and BMM. Adapted from [38,44].
Table 1. Comparative analysis of BCM, BIM, and BMM. Adapted from [38,44].
MechanismPrecipitate LocationConditionsOrganismsLevel of Organism ControlPrecipitated Minerals
BCMIntracellular, intercellular, extracellularControlled by cellular activitiesEukaryotesHighMagnetite, greigite, amorphous silica, calcite
BIMExtracellularReactive surfaces & metabolismProkaryotesModerateIron hydroxides, magnetite, manganese oxides, clays, amorphous silica, carbonates, phosphates, sulfates, sulfide minerals
BMMEPS matrixAlkalinity engine & organic matterNot requiredLowCarbonate minerals
Table 2. Comparative analysis of MICP and MCP.
Table 2. Comparative analysis of MICP and MCP.
Carbonate PrecipitationMechanismMicrobial InvolvementApplicationResearch TopicsAdvantagesDrawbacks
MICPBIMActiveIn situ 1 & ex situ 2Restoration of calcareous stones & construction materials, soil strengthening, selective plugging for oil recovery, bio-clogging, soil thermal conductivity, dust suppression, erosion control, liquefaction mitigation, wastewater treatment, bioremediation, CO2 sequestration [45]Wide range of applicable microorganisms, applicable to a wide range of environments, low costs, high CaCO3 conversion, short timeframes [45]Potential for harmful byproducts, bio-clogging at injection site, requires specific conditions
MCPBMMPassiveIn situ 1 & ex situ 2Wastewater treatment, oil recovery, biofilm barriers, bioremediation [46]Wide range of environments, adaptable to versatile environmental conditionsVariable efficacy for carbonate precipitation, slower rates of precipitation
1 In situ biostimulation, ex situ biostimulation and bioaugmentation [47]. 2 Material pre-treatment [35].
Figure 1. Biochemical precipitation mechanisms, including biologically controlled mechanisms (left), biologically induced mechanisms (right), and biologically mediated mechanisms (top). The red lines represent passive diffusion, blue lines represent active pumping, and green lines represent secretion. Adapted from [33,36,45].
Figure 1. Biochemical precipitation mechanisms, including biologically controlled mechanisms (left), biologically induced mechanisms (right), and biologically mediated mechanisms (top). The red lines represent passive diffusion, blue lines represent active pumping, and green lines represent secretion. Adapted from [33,36,45].
Ijms 26 02230 g001
It should be noted that in addition to bacterial carbonate precipitation, eukaryotes (i.e., coccolithophores and foraminifera) can precipitate carbonates to form shells or skeletons via BCM [44,48]. The formation of CaCO3 exoskeletons plays an important role in the carbon cycle, impacting CO2 flux in seawater and inorganic carbon transport in oceans and sediments [48]. Coccolithophores, specifically, are promising for global climate change because (i) they are phytoplankton (autotrophic plankton obtaining energy through photosynthesis); (ii) they produce dimethyl sulfide (DMS), creating albedo effects via formation of highly reflective clouds; (iii) they control CO2 influx into water via precipitation of CaCO3, which depletes dissolved bicarbonate (HCO3), increasing dissolved CO2 [44,48,49]. However, increased CO2 has been shown to decrease CaCO3 precipitation in marine phytoplankton [50], since increased atmospheric CO2 also increases carbonic acid (H2CO3), producing more HCO3 and H+ ions (Equation (2)), which dissolves CaCO3, decreasing pH [48]. Furthermore, there are numerous autotropic carbon-fixation mechanisms identified in archaea, including the Calvin cycle, reductive citric acid cycle, reductive acetyl-coenzyme A pathway, 3-hydroxypropionate bicycle, hydroxypropionate-hydroxybutyrate cycle, and dicarboxylate-hydroxybutyrate cycle [51].
C O 2 + H 2 O H 2 C O 3 H C O 3 + H + C O 3 2 + 2 H +

2.3. Microbial Carbonate Precipitation (MCP)

The precipitation of CaCO3 is dependent on the calcium (Ca2+) concentration, DIC, nucleation sites, and pH [27,52]. For the purpose of microbial precipitation, the principal role of bacteria is to create an alkaline environment via pH and DIC increase [52]. This can occur as a BMM (passive process) or BIM (active process) [26].
Biologically mediated CaCO3 precipitation occurs from the interaction of EPS and Ca2+ ions. As mentioned previously, the process does not require direct biological activity but is influenced by the organics associated with the cell wall and/or the EPS [26,34,37,53]. The microorganisms secrete natural polymers (polysaccharides, lipids, proteins, etc.), forming an organic matrix. These organic polymers favor heterogenous nucleation, leading to stabilization of new particles [37]. They can also act as nucleation sites [53,54]. Further, an increase in pH causes functional groups to deprotonate, causing exopolysaccharides produced by the microbial cell to have an overall negative charge and bind to metal(loid) ions [26].
The bacterial cell surface and biomass surfaces have an electronegative charge due to the presence of carboxyl, phosphoryl, amino, and sulfo groups [55]. The negative surface charge allows redox processes, adsorption, complexation, ion exchange, electrostatic attraction, and precipitation to immobilize metal(loid)s in situ [27,56]. For example, adsorption of positive divalent cations to the bacterial cell wall can lead to the precipitation of carbonates [27]. Initial adsorption of bacteria onto mineral surfaces is governed by hydrophobicity and electrostatic forces, while final attachment of bacteria to minerals is influenced by biofilm formation and secretion of exudates [57]. Therefore, uptake of cationic metal(loid)s can create a state of oversaturation in the microenvironment, leading to precipitation [58]. Further, both the cell wall and EPS have metal binding capacities; however, fate transport of these bonded metals is not known [59]. The bacterial cell wall provides nucleation sites for mineral deposition of biological precipitates [60,61]. This is in addition to exopolymers, biofilms, and inactive spores, which also provide sites for nucleation [26].
The function of EPS is important to the carbonate precipitation process. EPS are high-molecular-weight natural polymers (i.e., lipids, proteins, polysaccharides, DNA, etc.) secreted by autotrophic and heterotrophic microorganisms, responsible for the functionality and structural integrity of the biofilms [58,62]. The macromolecules (through dispersion forces, electrostatic interactions, and hydrogen bonds) create a gel-like substance around the cells, establishing a stable consortia of microorganisms [62]. Similar to the bacterial cell surface, EPS contains functional groups, including carboxyl, phosphoryl, amino, and hydroxyl groups [58,63,64]. These negative functional groups can attract positive divalent cations, thereby promoting precipitation of metal carbonate (MCO3) compounds through local alkalization or inhibiting precipitation by removing the free cations from solution and reducing saturation. If the latter transpires, MCO3 can be precipitated out of solution when EPS degrades and saturates solution with metal divalent cations [58,63]. The biochemical composition of EPS can affect the resulting mineralogy of CaCO3, altering the polymorph (aragonite, vaterite, calcite) [65]. It can also alter the crystal morphology of CaCO3 precipitates [66,67]. EPS is thought to influence the biofilm, cell adhesion, and capturing CaCO3 precipitates [54]. Both EPS and biofilm formation can reduce pore space, increase ductility, increase strength, reduce hydraulic conductivity, and reduce permeability [68].

2.4. Microbial-Induced Carbonate Precipitation (MICP)

MICP is a complex process involving numerous metabolic pathways. These metabolic pathways can be enzyme-driven, redox-driven, or photosynthesis-driven reactions (Figure 2) [28]. Equations (3) and (4) illustrate the governing equations for calcium carbonate precipitation induced by biological processes [69]. The Ca2+ ion can be interchanged by other divalent cations to precipitate other MCO3 compounds [45].
C a 2 + + C e l l C e l l C a 2 +
C e l l C a 2 + + C O 3 2 C e l l C a C O 3
Since SOC is considered energy limiting for microbial growth and carbonate mineralization [71], a nutrient broth (NB) is often added to supplement nutrient-deficient substrates for microbial growth. It is a complex concoction of chemicals and nutrients dependent on the desired microbial pathway. Nutrients often include carbon, nitrogen, and phosphorous, whereby an optimized ratio (C:N:P) is analyzed, typically approximately 100:10:1 [72]. The enzyme activity of urease, phosphatase, and dehydrogenase is involved in biogeochemical cycles of phosphate, nitrogen, and oxidation reduction of organic compounds and can illustrate the fertility of substrates [73].
Essential to MICP is a Ca2+ source. As mentioned above, Ca2+ concentration plays a vital role in CaCO3 precipitation. Typically, a calcium source is added to facilitate precipitation. The most common calcium source used in research is calcium chloride (CaCl2). The CaCl2 compound undergoes dissolution, and the Ca2+ ions precipitate CaCO3, while the chloride ions (Cl) form ammonium chloride (NH4Cl), shown in Equation (5) [27,74]. However, calcium acetate (C4H6CaO4) and calcium nitrate (Ca(NO3)2) are also used to facilitate carbonate precipitation [27]. Enhanced MICP is linked to higher concentrations of urea and CaCl2 [75]. Stoichiometric calculations are required to determine the amount of urea necessary to convert all Ca2+ to CaCO3 [76].
C l + H C O 3 + N H 3 N H 4 C l + C O 3 2
The interaction of calcium with the microbial cell is an intricate process. The microorganism is surrounded by a thin layer of water, and when subjected to a low Reynold’s number, protons (pH), DIC, and Ca2+ can concentrate in the microenvironment [52]. Ca2+ accumulates outside the microbial cell wall and is not likely utilized by the microbial metabolic processes [60]. McConnaughey and Whelan [77], Castanier [78], and Hammes and Verstraete [52], among other researchers, characterize the difference between “active precipitation” and “passive precipitation”, whereby the former is linked to ion transport and exchange (specifically Ca2+) through the cell membrane, and the latter encompasses precipitation via the metabolic pathways discussed below. Active precipitation is governed by calcium regulation via influxes and outflows (Figure 3). The transport mechanisms enabling active precipitation can be further classified as active or passive. Ca2+ influxes can be a passive transport mechanism based on the electrochemical gradient [52]. Concentrations of Ca2+ in the extracellular environment are typically 1000 times greater than in the intracellular environment due to low permeability of the cell envelope, high buffering capacities, and effective export mechanisms [79]. The passive transport mechanisms include antiporters (Ca2+/2H+, Ca2+/2Na+, etc.), protein-based channels, and non-proteinaceous channels. Active transport, however, includes Ca2+ transport against the electrochemical gradient using ATP-energy and ATP-dependent pumps [52].
The metabolism of calcium by the microorganism is a driving factor for CaCO3 precipitation. In the microenvironment, when Ca2+ concentrations and pH (low H+ proton concentration) are high in comparison to the microbial intracellular environment, the difference in electrochemical gradient will cause the Ca2+/2H+ antiporter to accumulate Ca2+ in the microbial cell and release H+ protons to the extracellular environment. Through active extrusion, the microorganism will then release Ca2+ through ATP-dependent calcium pumps and uptake H+ protons. This will create localized alkaline conditions and high Ca2+ concentrations ideal for precipitation. The metabolism of organic matter is required for ATP, which releases DIC to the extracellular environment in the form of CO2. The CO2 will undergo hydrolysis to form HCO3 and CO32− ions (Equation (6)), which will interact with the Ca2+ ions and precipitate CaCO3. This will impact the CaCO3 solubility product. As the soluble Ca2+ ions decrease and there is an increase in acidity, the conditions become favorable for bacterial proliferation [52].
C a 2 + + H C O 3 C a C O 3 + H +

2.4.1. Nitrogen Cycle

The nitrogen cycle plays a significant role in MICP. Three of the MICP metabolic pathways rely on the nitrogen cycle: denitrification, ammonification, and urea hydrolysis. In each of these scenarios, a pH increase (NH4+ and OH) in the presence of Ca2+ ions lead to CaCO3 precipitation. Shown in Equation (2), a decrease in H+ ions shifts the CO32−–HCO3 equilibrium to its CO32− form, inducing CaCO3 precipitation [78].
Denitrification (Equation (7)) utilizes nitrate-reducing bacteria (NRB) to facilitate precipitation [58,80]. The dissimilatory reduction of nitrate (NO3) increases pH through the consumption of NO3 and the generation of OH ions [81]. Under anaerobic or hypoxic conditions, NO3 acts as the electron acceptor producing inorganic carbon in the form of CO2 [58,81]. Nitrogen (N2) gas is an end product of dissimilatory nitrate reduction; however, intermediates include nitrite (NO2), nitric oxide (NO), and nitrous oxide (N2O) [82]. Toxic intermediates (NO2 and N2O) can accumulate if the involved enzymes are inhibited [58,83]. There are four enzymes involved in the denitrification process: nitrate reductase (Nar), nitrite reductase (Nir), nitric oxide reductase (Nor), and nitrous oxide reductase (Nos) [82]. The localization, lifetime, regulatory mechanisms, kinetics, and sensitivity to inhibitors are different for each of these enzymes, which can lead to incomplete denitrification at any of the reduction steps [83]. For example, nitrate or calcium overloading can cause inhibition, yielding the accumulation of toxic intermediates [83]. Synthesis of these enzymes is dependent on oxygen (O2) concentration, pH, and temperature [84]. Evidence suggests nitrate plays a role in carbon-fixing pathways during carbon sequestration via revegetation, whereby nitrate directly impacts soil labile organic carbon and indirectly influences carbon-fixing microorganisms [85]. Furthermore, microbial genes identified in soils are involved in the nitrogen cycle and carbon fixation [86].
N O 3 + 5 4 C H 2 O 1 2 N 2 + 5 4 C O 2 + 3 4 H 2 O + O H
Ammonification (Equation (8)) utilizes amino acids to produce NH3 and CO32− via myxobacteria to induce precipitation [58]. Myxococcus xanthus, for example, is shown to induce precipitation of calcite and vaterite crystals [87,88,89]. This occurs under aerobic conditions with gaseous or dissolved oxygen and organic matter. These heterotrophic microorganisms use amino acids as an energy source and contribute to the degradation of organic matter [58,78]. The hydrolysis of NH3 produces hydroxide ions (OH), creating a pH increase and leading to local supersaturation around the microbial cell, favoring precipitation [58]. Also formed from NH3 hydrolysis is the NH4+ byproduct, often present in an aqueous state. NH4+ can easily convert into NO3 and NO2 [90], which can cause the accumulation of toxic nitrogen species in the environment. In addition, NH4+ in surface water can promote toxic algal bloom growth, impacting fish, flora, and fauna [91].
A m i n o   A c i d + O 2 N H 3 + C O 2 + H 2 O
Urea hydrolysis involves the degradation of urea via ureolytic bacteria. The process consists of three main stages: (i) urea hydrolysis; (ii) pH increase; and (iii) cementation [69]. Ureolytic bacteria required for urea hydrolysis have a direct impact on the concentration of DIC (cell respiration and the decomposition of urea) and pH within the environment [58]. As part of the urea hydrolysis reaction (Equations (9) and (10)), 1 mol of urea (CO(NH2)2) via hydrolysis produces 1 mol carbamic acid (NH2COOH) and 1 mol NH3, where 1 mol NH2COOH undergoes spontaneous hydrolysis to produce 1 mol NH3 and 1 mol carbonic acid (H2CO3) [26,27,92]. This occurs via secretion of the enzyme urease (urea amidohydrolase; E.C. 3.5.1.5; nickel-containing metalloenzyme), which acts as a biological catalyst [93,94]. The enzyme speeds up the chemical reaction by lowering the activation energy via low-energy enzyme–substrate (i.e., urea) complexes. The bacteria use urease to hydrolyze CO(NH2)2 (added to the NB; NBU) by increasing ambient pH and using CO(NH2)2 as a nitrogen and energy source [60,93]. The NH3 (Equation (11)) produced from enzymatic urea hydrolysis will again undergo hydrolysis to form NH4+ and OH, which increases pH [26,27,92]. The pH increases via urease can create a localized alkaline state in the microenvironment around the microbial cell, leading to CaCO3 precipitation on or around the cell wall [60,74]. Urease production and subsequent CaCO3 precipitation are impacted by the temperature, pH, concentration of CO(NH2)2, concentration of NH3, carbon source, and incubation period [95].
C O ( N H 2 ) 2 + H 2 O N H 2 C O O H + N H 3
N H 2 C O O H + H 2 O N H 3 + H 2 C O 3
2 N H 3 + 2 H 2 O 2 N H 4 + + 2 O H
Through the precipitation of CaCO3 is the development of cementation, which pertains to the large-scale precipitation in-between solid particles forming a biocement matrix [35,45]. It involves the dissociation of a calcium source into soluble Ca2+ ions [81], and an increase in CO3 ions to reach the supersaturation state inducing CaCO3 precipitation [69]. Again, NH4+ is produced as a byproduct. It should be noted that urea is very stable, and the purely chemical breakdown is independent of the pH between 2 and 12 [96]. The non-enzymatic process decomposes urea via elimination of NH3 (half-life of 33 years at 25 °C) or spontaneous hydrolysis (half-life of 520 years at 25 °C) [97]. However, application of urease drastically increases the rate of reaction to a half-life in the microsecond range [97,98].
Several methods have been explored to reduce or remove harmful NH4+ byproduct from nitrogen-driven MICP pathways. These include flushing and extracting NH4+ with geophysical setups, electrokinetic retention of NH4+ in the cathode chamber of an electrokinetic cell, NH4+ precipitation via additives, and utilization of alternative metabolic pathways [99]. An alternative metabolic pathway is iron reduction utilizing iron-reducing bacteria [80,100]. Ferric iron (Fe3+) acts as an electron acceptor in the presence of a carbon source, reducing to ferrous iron (Fe2+) and CO2 (Equation (12)) [80]. However, mineral precipitates are unstable and easily impacted by other ions [47]. Current research on this metabolic pathway is limited, although application of ureolytic MICP to iron-based substrates is emerging [101,102]. Iron-reducing bacteria have shown an impact in the complex coupling of Fe and C affecting carbon sequestration in paddy soils [103]. Furthermore, Fe can trap SOC via adsorption, coprecipitation [104], whereby ~21.5% of SOC is bound to reactive Fe phases in sediments [105].
C H 2 O + 4 F e 3 + + H 2 O 4 F e 2 + + C O 2 + 4 H +
The efficacy of the urease enzyme can be impacted by nickel. The urease enzyme is composed of structural genes and accessory genes in operons and clusters [26,106]. Inactive urease (apo-urease) has structural genes (ureA, ureB, and ureC) requiring accessory genes (ureD, ureF, ureG, and ureE) for activation. Activation involves CO2 uptake for lysine carbamylation, hydrolysis of guanosine triphosphate (GTP), and Ni2+ delivery to its active site [107]. As mentioned preciously, nickel is incorporated in the active center of urease [26] and contains two nickel ions (Ni2+) bridged by a hydroxyl group and a carbamylated lysine [108]. The ureE gene, specifically, is responsible for delivering Ni2+ to the active site, leading to fully active urease (holo-urease) and subsequent urea hydrolysis [107]. A mobile flap (from the helix-turn-helix motif) covers the active site, restricting access [109]. As shown in Figure 4, ureolysis occurs when the flap is open and urea enters the active site, replacing water molecules bound to Ni2+ ions. The C-O bonds in urea are polarized and undergo nucleophilic attack because of the highly electrophilic Ni ions with the bridging OH [107,109,110]. The NH2 is protonated by the bridging Ni OH (or His320, Ala167, Ala363, Cys319, His219, G277) [110], and the C-N bond is broken, releasing NH3 [107,109]. The carbamate (CH2NO2) remaining decomposes further (Equation (13)), and all products are released with the flap opening [107]. Furthermore, urease inhibition is attributed to Ni2+ binding, which leads to a loss in urease and catalytic activity. This includes impacts to the direct binding of Ni2+ at the active urease sites; covalent modifications that cover the Ni2+ center; and metal ion chelators that sequester Ni2+, thereby inhibiting the formation of the Ni2+ center [108]. Ni2+ in the form of nickel chloride (NiCl2) and nickel nitrate (Ni(NO3)2) has been added to the NB to enhance CaCO3 precipitation [111,112]. In addition to ureolysis, nickel is involved in hydrogen metabolism and methane biogenesis, and acts as an essential nutrient to microorganisms [109].
N H 2 C O O + 2 H 2 O N H 4 + + H C O 3 + O H
Application of ureolytic MICP for carbon sequestration showed efficacy ≤86.4%, dependent on the bacterial community structure and pH [113]. Higher-headspace CO2 uptake was shown with Sporosarcina, Sphingobacterium, Stenotrophomonas, Acinetobacter, and Elizabethkingia species [113]. An increase ≤ 148.9% in CO2 uptake through calcification can be shown in optimal urea growth media [33]. Conversely, ureolytic bacterial growth utilizing Bacillus megaterium demonstrated comparable quantities of precipitated CaCO3 with 99.5% pure CO2 influx to that of 2% NBU [114].

2.4.2. Sulfur Cycle

The sulfur cycle also plays an interesting role in MICP (Equation (14)). In sulfate (SO42−)-rich environments, sulfate-reducing bacteria (SRB) facilitate either dissimilatory or assimilatory sulfate reduction, producing hydrogen sulfide (H2S) or organic sulfur (S), respectively [115,116]. This reaction transforms organic carbon in the form of an energy source to HCO3, and will release OH and increase alkalinity, leading to a supersaturation state and therefore the likelihood of CO32− precipitation [58]. SRB (comprising 87% Halanaerobiaceae, Halobacteroidaceae, Enterobacteriaceae) have been enriched from sediment samples for carbon sequestration, whereby ~53% of precipitated carbonate minerals are derived from CO2 headspace [117]. In this study, headspace pressure played an integral role in carbonate precipitation at ≤ 14.7 psi [117]. The SRB can also degrade EPS, releasing trapped Ca2+ into the environment, leading to CaCO3 precipitation [58]. Further, the H2S produced during the reaction may subsequently degas, increasing pH, increasing precipitation, or being utilized by bacteria [78]. However, the H2S can be highly toxic [81], and if not degassed or if unused by bacteria, it can cause pH to decrease and inhibit precipitation [78]. The process is most prominent under anaerobic or anoxic conditions rich in organic matter [26,78]. Sulfate reduction contributes ~36–50% carbon mineralization in anaerobic wetlands [118] and plays a significant role in stromatolite formation [119].
S O 4 2 + 2 C H 2 O H 2 S + 2 C O 2 + 2 O H
Methanogens also utilize SO42− to induce methane oxidation [58]. Under anaerobic conditions, methane oxidation favors carbonate precipitation (Equation (15)), while aerobic conditions increase alkalinity, favoring carbonate dissolution (Equation (16)) [58,120]. Both scenarios can facilitate carbonate precipitation in the presence of a divalent cation source (i.e., Ca, Fe, Mg, Mn, Ba) [121]. However, interesting is the removal of CH4 emissions and its application as a methane sink offsetting GHG emissions [122]. The process transforms CH4 to a less toxic form and locks carbon as mineral precipitates. Furthermore, in situ anaerobic carbonate precipitation via methane oxidation can be subdivided into sulfate-dependent precipitation in shallow sediments or marine silicate weathering in deep sediments [121]. Unlike the other metabolic pathways that are solely heterotrophic, methane oxidation can be both an autotrophic [26,123] or heterotrophic process [124]. It can also use alternative electron sources to SO42−, including NO3, NO2, Fe, Mn, and humic acid [120]. While methane oxidation in terms of MICP is often neglected in experimental research, a recent study has identified methanogenesis as a metabolic pathway in activated anaerobic sludge [99].
S O 4 2 + C H 4 H S + H C O 3 + H 2 O
2 O 2 + C H 4 C O 2 + 2 H 2 O

2.4.3. Photosynthesis

Purely autotrophic pathways (i.e., non-methylotrophic methanogenesis, anoxygenic photosynthesis, and oxygenic photosynthesis) utilize gaseous or dissolved CO2 from the atmosphere, respiration, or fermentation [78]. This CO2 acts as their carbon source to produce organic matter and, under Ca2+ rich environments, can favor CaCO3 precipitation [78]. Photolithoautotrophs, specifically, cyanobacteria, have been studied for their capacity for carbonate biomineralization [120], with their higher affinity for environment CO2 due to photosynthesis and CO2 fixation by the Calvin cycle [125]. Approximately 70% of carbonate rock in the history of Earth is contributed to cyanobacteria [126]. Cyanobacteria carbonate mineralization occurs in four steps: (i) CO2/HCO3 uptake for photosynthesis; (ii) OH release; (iii) OH and HCO3 reaction forming CO32−; (iv) carbonate precipitation [127]. The CO2 (Equation (17)) enters the cell wall via diffusion or a symporter, where CO2 produces organic matter utilized by the microorganisms [58]. HCO3 (Equation (18)) is transported from the extracellular environment into the cell membrane, which dissociates into CO2 and OH [26]. HCO3 is the predominant form of inorganic carbon transported into the cell [127]. The OH ion is released from the cell to the extracellular environment, increasing pH, which again favors CaCO3 precipitation in Ca2+-rich environments [26,58]. The equilibrium reached between HCO3 into the cell and the efflux of OH from the cell causes the alkalinization in the microenvironment around the cell [127]. CaCO3 nucleation occurs on the sticky cell walls of cyanobacteria, which aid in binding [128]. Furthermore, calcium metabolism can store Ca2+ within the cell membrane, precipitating CaCO3 intracellularly, or can release Ca2+ through the Ca2+/2H+ antiporter, precipitating CaCO3 extracellularly [58]. This process has been investigated more frequently due to the nature of its less harmful byproduct (CH2O). The absence of NH4+ or H2S make this pathway desirable. Naturally occurring photosynthesis-driven carbonate mineralization is shown in karstic environments to offset mine-related GHG emissions by ~20% [129,130].
C O 2 + 2 H 2 O C H 2 O + O 2
H C O 3 C O 2 + O H
Carbonic anhydrase (CA; EC 4.2.1.1; zinc-containing metalloenzyme) is an enzyme-driven metabolic pathway essential for carbon sequestration [34,93]. The enzyme catalyzes the reversible hydration of CO2 (Equation (2)) and plays a role in pH regulation [120]. The enzyme is very complex, containing five distinct classes (α, β, γ, δ, and ε; evolutionarily independent) [131], and α-CA contains 15 isozymes [132]. The enzyme can be intracellular, intra-organellar, periplasmic, or extracellular [120]. Intracellular CA can possess a carbon-concentrating mechanism (CCM) to capture and sequester CO2 [131,133,134]. Inorganic-carbon-concentrating mechanisms (Figure 5) can occur from (i) diffusion or active transport of carbon (CO2 or HCO3) across the cytoplasmic membrane via energy-dependent transporters; (ii) CA conversion to HCO3, which accumulates in the carboxysome; (iii) conversion of HCO3 back to CO2 by CA, which concentrates and fixes elevated concentrations of CO2 in the Rubisco [125,133]. The CCM is thought to have evolved from the decline of atmospheric CO2 and increase in O2 in the Phanerozoic era, triggering oxygenic photosynthesis in cyanobacteria [135]. The CCM genes (ccmK, ccmL, ccmM, ccmN, and ccmO) enable growth at low pCO2 for assembly in the carboxysome [125]. The carboxysome (sub-cellular compartment encapsulating Rubisco and CA) is a primary component of CCM [136]. There exists a differentiation between α-cyanobacteria and β-cyanobacteria based on the type of carboxysome and Rubisco [135]. The carboxysome is separated into α-carboxysome and β-carboxysome, both of which limit CO2 leaching; reduce the risk of photorespiration; and enhance carboxylase, the activity of Rubisco [136]. The CA enzyme is specific to the type of carboxysome: α-carboxysome requires β-CA (CsoSCA), and β-carboxysome requires β-CA (CcaA) and γ-CA (CcmM) [136,137], but β-CA has a direct involvement in CCM of β-cyanobacteria [135]. β-cyanobacteria also contain non-carboxysomal CA localized in the cell membrane or periplasmic space: α-CA (EcaA) and β-CA (EcaB) [136]. Cyanobacteria have been identified in hot/cold, alkaline/acidic, marine, freshwater, saline, terrestrial, and symbiotic environments [138]. The CCM of α-cyanobacteria and β-cyanobacteria is dependent on their environmental conditions (i.e., pH, carbon content, salinity, temperatures, oxygen content, light, wet/dry conditions) [138]. However, pH is most prominent since it is linked to carbon speciation (i.e., H2CO3, CO2, HCO3, CO32−) [136].
However, α-CA are localized in the periplasmic or extracellular space, and it is hypothesized that they are able to convert diffused CO2 into HCO3 for bacterial metabolism [139]. Extracellular α-CA has been identified from prokaryote Pseudomonas fragi [140], Bacillus sp. [131], cyanobacteria Microcoleus chlthonoplastes [141], Bacillus mucilaginosus [142]. Both intracellular and extracellular CA could be detected in soil bacteria, whereby some CA was absorbed by soil [143]. The extracellular CA likely stabilizes the pericellular pH and induces carbonate precipitation [141] at or near the bacterial cell wall.
While nickel is at the core of urease, CA contains a zinc core [144]. The zinc plays a vital role in the CA activity [145]. There are several genetically distinct forms of CA, each containing a catalytically obligatory zinc ion (Zn2+) [146]. The hydrogen bonding network stabilizes the electrostatic environment of zinc, impacting catalytic efficacy [147]. The structure of CA (Figure 6) and orientation of CO2 enhance the likelihood of nucleophilic arrack of the Zn2+-bound water molecule. This leads to HCO3 formation, which is subsequently replaced by a water molecule, releasing HCO3 from the active site [148]. The addition of zinc as zinc sulfate (ZnSO4) can have a positive effect on CA activity [144]. However, studies show that the zinc ion in the enzyme is firmly bound to the protein, forming a stable metal–protein complex, and therefore, ion exchange with Zn2+ in solution is unlikely [149]. CA inhibition can occur via bonding to the zinc-coordinated water molecule/hydroxide ion [134]. This can occur via metal complexing anions and substitution of the non-protein zinc ligand [148].
Figure 5. A generic model of the CCM of cyanobacteria showing accumulation of HCO3 in the cytosol, Rubisco-containing carboxysome, and CA. Intracellular pH is buffered via Ca2+/H+ antiporter, which alone releases OH in the extracellular environment and increases pH, favoring extracellular CaCO3 precipitation. Adapted from [124,138,140,142].
Figure 5. A generic model of the CCM of cyanobacteria showing accumulation of HCO3 in the cytosol, Rubisco-containing carboxysome, and CA. Intracellular pH is buffered via Ca2+/H+ antiporter, which alone releases OH in the extracellular environment and increases pH, favoring extracellular CaCO3 precipitation. Adapted from [124,138,140,142].
Ijms 26 02230 g005
The enzyme-driven pathways (urease and CA) can be independent processes, but they can also work synergistically to facilitate carbon sequestration [54,93,152]. CO2 dissolution can decrease pH due to proton enrichment [21,153]. However, there will also be a pH increase from NH4+ from urea hydrolysis [154], which maintains the alkaline state required for precipitation [93]. Further, the urease enzyme has nickel incorporated in the active center, and an increase in CO2 (regulated by CA) is shown to generate ligands for nickel binding, which is essential to urease activity [26,155]. The nickel core is dependent on CO2/HCO3 metabolism [93]. However, it should be noted that carbon sequestration can occur independently as a purely chemical reaction [33], whereby CO2 is dissolved in water, converted to CO32−, and then precipitated as CaCO3 without biological interference.

3. Carbon Sequestration

Carbon sequestration methods are characterized as direct (i.e., involved in reduction in CO2 emissions by sequestering inorganic carbon prior to atmospheric release) or indirect (i.e., reliance on natural carbon sinks) [41]. The direct methods require CCS techniques for CO2 removal, which includes absorption (chemical, physical), adsorption (adsorber beds, regeneration methods), cryogenics, membranes (gas separation, gas absorption, ceramic-based systems), and microbial/algal systems [41]. Mechanisms for carbon sequestration are outlined in Figure 7.
As shown in Figure 7, there are 4 major carbon pools responsible for indirect carbon sequestration. A carbon pool refers to a system that can accumulate or release carbon [157], and these include oceanic, geologic (fossil fuels i.e., coal, oil, gas, peat), terrestrial (pedologic (i.e., SOC, soil inorganic carbon (SIC)) and biotic (i.e., vegetation), and atmospheric pools [40]. A major factor in the global warming crisis is the depletion of fossil fuel carbon pools and the anthropogenic release of GHGs back into the atmosphere. The majority of GHGs are released by the combustion of fossil fuels for energy and transportation. Current methods of carbon sequestration are typically defined as abiotic (oceanic injection, geologic injection, mineral carbonation) or biotic (oceanic sequestration, terrestrial sequestration, mineral carbonation) processes [40].
Oceanic carbon sequestration can be both an abiotic and biotic process. These techniques rely on the solubility pump and autotropic mechanisms [41]. Methods can stimulate growth of autotrophic organisms (phytoplankton, microalgae, macroalgae, and cyanobacteria) on the ocean surface to enhance photosynthesis to remove atmospheric CO2 [41]. The methods can utilize DOC for: (i) photosynthesis; (ii) remineralization; (iii) assimilation by microorganisms [158]. The trapped carbon via biological carbon pumps (i.e., gravitational settling, ocean mixing and animal migrations) to mobilize the organic matter downward for burial at the ocean bottom [159,160]. Alternatively, CO2 can be injected in liquified phase directly into the deep ocean (> 1 km) to form CO2 hydrate for permanent storage [161].
Deep geologic injection, as named, involves deep geologic injection of supercritical CO2 into porous aquifers (i.e., coal seams, oil beds, deep saline aquifers) capped by low permeability rock [162]. The CO2 can be trapped by: (i) mineral trapping through precipitation of carbonate minerals; (ii) geologic trapping by physical containment in geologic features; (iii) solubility trapping via dissolution in liquid; (iv) hydrodynamic trapping from CO2 and liquid viscosity differences; (v) capillary trapping due to capillary forces; (vi) sorption of CO2 on the materials surface [21]; (vii) formation trapping by reduced geologic permeability to reduce CO2 leakage [53]. In saline aquifers, the CO2 injected in supercritical state can be sequestered hydrodynamically by reacting with dissolved salts forming carbonate minerals, by an additive of lower density. The lower viscosity solution displaces brine, which creates a multiphase (gas-like and aqueous) environment [40]. Sedimentary basins are well suited for CCS via deep geologic injection due to the high pore volume and connectivity [163].

3.1. Mineral Carbonation and Carbon Sequestration

Mineral carbonation is significant to both abiotic and biotic carbon sequestration methods. The method can be naturally occurring mimicking the natural weathering process of alkaline silicates (Equations (19)–(21)) [164]. The process dissolves atmospheric CO2 in rainwater to process weak carbonic acid (H2CO3), which is slightly acidic causing metal ions to leach from natural alkaline silicates neutralizing their mineral alkalinity and precipitating of carbonates [164]. The process precipitates geologically, geochemically and thermodynamically stable carbonate precipitates [40,164,165]. These precipitates are low solubility [166], and would require acidic conditions or high temperatures (~900 °C) to release CO2 from the mineralized precipitate [167]. Mineral carbonation is therefore considered permanent solution for carbon sequestration of atmospheric CO2 [168]. Carbonation is impacted by: solid to liquid ratio, particle size, temperature, ion transport mechanisms [166], pH, crystal ageing, agitation, and impurities [169]. It should be noted that environments with high pH leachate from weathering with high CaCO3 precipitation can: (i) smother benthic ecosystems; (ii) damage littoral aquatic habitats; (iii) reduce light penetration to benthic primary producers; and (iv) harm fish populations [170].
C a S i O 3 ( s ) + 2 C O 2 ( a q ) + H 2 O ( l ) C a 2 + ( a q ) + 2 H C O 3 ( a q ) + S i O 2 ( s )
M g S i O 4 ( s ) + 4 C O 2 ( a q ) + 2 H 2 O ( l ) 2 M g 2 + ( a q ) + 4 H C O 3 ( a q ) + S i O 2 ( s )
( M g 2 + , C a 2 + ) a q + C O 3 2 a q ( M g , C a ) C O 3 ( a q ) ( M g , C a ) C O 3 ( s )
Accelerated carbonation is a process used to replicate the natural weathering process. It speeds up the process by utilizing high-purity CO2, which reacts with alkaline materials in the presence of moisture to precipitate carbonates within minutes or hours [171]. There are two main categorizes of processes: the direct method (single reaction step) and the indirect methods (alkaline metal ions are extracted prior to carbonate precipitation in a multi-step process) [164,165,172]. A primary advantage of indirect processes is the production of pure CaCO3 (or MgCO3) without impurities (i.e., silica) [173]. The process can further be classified as in-situ or ex-situ approaches [174,175]. The former injects CO2 directly into the porous material to react with the host rock, whereas the latter uses industrial chemical processes to carbonate natural minerals and industrial waste in treatment plants [174]. Interestingly, the use of alkaline waste originates from industrial and mining activities as host rock for carbon sequestration [165,166,172,176], whose operations and feedstock are often located near point-source GHG emission sources [175]. The operations are promising for storage of atmospheric CO2 as carbonate precipitates and offsetting CO2 emissions from high-GHG producers. Natural CO2 sequestration has been demonstrated in chrysotile mine tailings in Clinton Creek, YT and Cassiar, BC, whereby the process is accelerated by the increased surface area from milling [177]. Researchers have also incorporated CA into accelerated mineral carbonation of alkaline brucite (Mg(OH)2) to overcome the carbonation rate-limiting supply of CO2, demonstrating acceleration of 240% over controls [178].
Direct mineral carbonation methods (Table 3) can be gas-solid carbonation or aqueous carbonation (gas-liquid or gas-liquid-solid) [164]. The operation is simple relying on an input of CO2 to facilitate precipitation of carbonate minerals. Direct precipitation can occur under dry or moist conditions [164]. However optimal CO2 sequestration often requires a degree of moisture [179,180,181,182]. The particle size also plays an important role, whereby smaller particle sizes are preferable [180,183]. Mechanical pretreatment (crushing and grinding) can be used to reduce particle size <300 μm, destroying the mineral lattice and increasing surface area for the reaction [175]. Researchers are also studying thermal pre-treatment, NETL derived processes, brine-based processes, and organic acid direct processes [175]. Although these processes can be considered indirect stepwise gas-solid methods [184]. Fagerlund et al. [185], for example, are studying stepwise carbonation of serpentinite (Equations (22) and (23)) at Åbo Akademi University, whereby magnesium ions are released from Mg3Si2O5(OH)4 through heat (400–500 °C) and ammonium sulphate ((NH4)2SO4) induce precipitation of magnesium sulfate (MgSO4), which afterwards is used to precipitate magnesium hydroxide (Mg(OH)2) via an aqueous ammonium hydroxide solution (NH4OH). The Mg(OH)2 precipitate is then used for carbonation with CO2 injection to precipitate MgCO3.
M g 3 S i 2 O 5 ( O H ) 4 + 3 ( N H 4 ) 2 S O 4 3 M g S O 4 + 2 S i O 2 + 5 H 2 O ( g ) + 6 N H 3 ( g )
M g S O 4 + 2 N H 4 O H ( a q ) ( N H 4 ) 2 S O 4 ( a q ) + M g ( O H ) 2
Alternatively, the direct aqueous method relies on the reaction of CO2 and water (Equation (2)) to form HCO3 and a proton (H+), which releases the divalent cation from the mineral (Equation (24)) to precipitate carbonate (Equation (25)) [199]. The reactions can be accelerated more via additives (i.e., bicarbonate/salts, acids, or chemical activators) or pretreatment methods (comminution, magnetic separation, heat treatment), which alter reaction conditions and modify solution chemistry to increase reaction rates to increase carbonate precipitation [172].
M g S i O 4 + 4 H + 2 M g 2 + + S i O 2 + 2 H 2 O
M g 2 + + 2 H C O 3 M g C O 3 + H +
Indirect mineral carbonation typically utilizes stepwise gas-solid, pH swing, or chemically enhanced mechanisms [184]. In the former, acid addition is utilized to enforce metal separation and a base additive induces aqueous carbonation [164,184]. The latter, chemically enhanced mechanisms, can include: HCl extraction, the molten salt process, other acid extractions (acetic acid (CH3COOH), sulfuric acid (H2SO4), nitric acid (HNO3), and formic acid (HCOOH)), bioleaching, ammonia extraction, and caustic extraction [172]. Acid extraction merely extracts the desirable metals from mineral prior to aqueous carbonation [164], causing potentially unfavorable conditions for precipitation (i.e., acidic conditions cause low CO2 dissolution and the low pH inhibits precipitation) [172]. HCl extraction, for example, showed no precipitated CaCO3 due to the rapid pH decline causing acidic conditions inhibition precipitation [200]. The pH swing process was developed by Park and Fan [201] to overcome these limitations. Wang and Maroto-Valer [202] have developed an amended pH swing method for serpentine incorporating the following steps: (i) CO2 is captured and reacts with NH3 to form NH4HCO3 intermediary; (ii) mineral dissolution via NH4HSO4 additive producing MgSO4; (iii) NH4OH is added to neutralize pH and remove impurities; (iv) MgSO4 reacts with NH4HCO3 at mild temperature to form Mg(HCO3)2 converting in the presence of water to MgCO3; (v) recovery of (NH4)2CO3 from carbonation via evaporation and heating to produce NH4HSO4 and NH3 for reuse.
Bioleaching utilizes byproducts (i.e., production of organic acids, chelating and complexing compounds) excreted by microorganisms to extract metals from minerals [203]. The humic and organic acids, inorganic acids, and chelating agents can free nutrients enhancing physical and chemical weathering [204]. These steps can be categorized by direct (Equation (26)) or indirect (Equation (27)) bacterial leaching, whereby both equations show metal sulfide (MS) oxidation into metal sulfate (MSO4) [203]. In Equation (26), bacterial enzyme activity catalyzes the mineral sulfate oxidation through direct physical contact between the bacterial cell and the mineral sulfide surface [203]. Metal sulfides include: covellite (CuS), chalcocite (Cu2S), sphalerite (ZnS), galena (PbS), molybdenite (MoS2), stibnite (Sb2S3), cobaltite (CoS), millerite (NiS), and pyrite (FeS2) [203]. The indirect oxidation in Equation (27), is generated by a catalytic function of a lixiviant which chemically oxidizes the sulfides [203]. In both scenarios, potentially acid generating substances (i.e., comprising sulfides or elemental sulfur) provide food for bacteria, which generated sulfuric acid as a by-product of metabolism, which leaches metals from minerals [172]. Furthermore, autotrophic bacteria (i.e., chemolithoautotrophic) can fix carbon biologically through the process by utilizing inorganic, atmospheric CO2 instead of organic carbon for new cell synthesis [172,203]. Factors influencing bioleaching include nutrients, O2 and CO2 content, pH (acidic conditions), temperature, mineral substrate (dependent on mineralogical composition and particle size), heavy metals, surfactants and organic extractants (decrease surface tension and mass transfer or oxygen) [203]. Chiang et al. [205] attributed bioleaching to organic acid production, specifically gluconic acid (C6H12O7), and microbial exopolysaccharides (EPSs). Similar to other mineral carbonation techniques, the released metal ions are available for carbonate precipitation. Bioleaching of ultramafic mine tailings can be used at tailings storage facilities [206]. Chrysotile tailing indicate acid mine drainage environments with microbial catalyst from Acidithiobacillus sp. is promising for MgCO3 precipitation with atmospheric CO2 resulting in 316 kt Mg leached/10 Mt tailings (458 kt CO2 sequestered/year) [207]. Argon Oxygen Decarburization slag showed a decline in primary phase (dicalcium-silicate, bredigite, and periclase) and an increase in secondary phases (merwinite and calcite), specifically a 3.1 wt % increase in CaCO3 with B. mucilaginous bacterial species [205].
M S + 2 O 2 B a c t e r i a M S O 4
M S + F e 2 ( S O 2 ) 3 M S O 4 + 2 F e S O 2 + S 0

3.2. Advancements to Mineral Carbonation for Carbon Sequestration

Microbial carbon mineralization has a unique opportunity to utilize novel biochemical mechanisms for carbon sequestration. The process can sequester inorganic carbon as a CCS technique via carbonate precipitation, but also mitigate CO2 released to the atmosphere therefore reducing atmospheric CO2 levels to aid the impacts of climate change [208]. The MICP enzyme-driven reactions speed up chemical reactions to enhance the rate of reaction and therefore conversion to CaCO3, optimizing storage of inorganic CO2. The synergistic role of CA and urease enzymes work to hydrate atmospheric CO2 while inducing an alkaline state for biocalcification [93]. Again, alkaline material is a preferential substrate for carbon sequestration [165,166,172,176], therefore halotolerant and alkalophilic bacteria are required for biocalcification. Ureolytic bacteria offer a preferable metabolic pathway for MICP, since microorganisms have shown capable of withstanding unfavorable conditions for bioremediation [154,209,210], and operate under high pH values and high concentrations of inorganic salts (i.e., CaCl2) [211]. Montmorillonite-coupled MICP in cyanide tailings showed up to a 1.33 increase in precipitation and up to 34.55% CO2 capture [212]. Other natural environments that supply cations for carbonation include: evaporate deposits, saline aquifers, waste brines, wastes from oil extraction, seawater [213]. MICP has shown success in precipitating carbonate in mine waste [73,154,214,215,216], concrete and building materials [217,218,219,220], coastal and marine environments [38], and agriculture and soil [47,221,222,223], all of which indirectly demonstrate the promising potential for carbon sequestration utilizing various substrates. Table 4 summarizes research investigating MICP as a method for carbon sequestration.
In addition to trapping CO2 as mineral precipitates, there exists addition mechanisms in which CO2 can be trapped and therefore stored. MCP and MICP have been utilized to aid CCS methods. During deep geologic injection the CO2 remains in supercritical state resulting in a distinct phases separate from formation water or brine, which is less dense and viscous permitting CO2 leakage [162]. Ureolysis and biofilm formation have been used to enhance mineral trapping, solubility trapping, and formation trapping of supercritical CO2 for geologic carbon sequestration [46,53,153,162,233]. Mineral trapping by precipitation of stable carbonates in deep geologic structures can store carbon but also reduce structural permeability to mitigate CO2 leakage [234]. Transition-state calcite and siderite have formed in deep saline aquifers by indigenous microorganisms [153]. The formation of biofilms in high-pressure pore spaces can decrease permeability by >95% trapping gaseous CO2 [46]. Furthermore, a pH increase generated by urea hydrolysis can increase DIC thereby lowering CO2 gas in the headspace for solubility trapping [233].
While carbonate precipitation can directly trap and store atmospheric CO2, it can also reduce GHG emissions released to the atmosphere. The cement industry is notorious for its significant contribution to the release of GHG emissions, accounting for 7% of the global GHG emissions and 1.5% (11.2 Mt in 2019) of Canadian emissions [235]. The production of cement (Figure 8) includes mineral extraction of raw materials, mineral processing, raw meal production, clinker formation, cement production and transport [236]. The process produces significant GHG emissions by: (i) calcination reactions (i.e., clinker process; Equation (28)); (ii) combustion of carbon heavy materials (i.e., coal, fuel, natural gas, petroleum coke, etc.); (iii) high energy requirement (2% global energy consumption); (iv) scale of production; (v) material treatment (grinding, mixing, additives); (vi) transport [237]. Self-healing bioconcrete, for example, uses autogenic (chemical reaction precipitating CaCO3 from cement hydration) and autonomic processes (application of encapsulation or a continuous vascular system to distribute a healing agent to precipitate CaCO3; can include MICP) as advancements to concrete repair [219,220]. MCP and MICP can be used as an alternative method for soil and concrete strengthening and restoration of calcareous stones and construction materials [45], reducing the usage of heavy GHG-producing cements. By precipitating CaCO3 or other MCO3 compounds in the pore space, cracks or fissures, MCP and MICP creates a clogging effect which strengthens the material and reduces compressibility [223]. This is in conjunction with the formation of a biocement matrix, whereby “preferential” and “uniform” distribution of precipitates at particle-particle contacts and precipitation around solid particles respectively, which improves engineering properties [223]. Application of MICP in cementitious materials for the construction industry showed a decrease in CO2 (3800 ppm to 820 ppm) with precipitated calcite and vaterite crystals through recombinantly produced CA MICP [225]. This shows MICP ability for carbon sequestration and carbon negative cementitious materials [225].
C a C O 3 C a O + C O 2
Finally, MCP and MICP can treat environmental disasters caused by climate change. Increased drought and more severe storms are widely accepted effects of climate change. Earthquake-induced liquefaction causes soil to behave fluidlike because of increases in pore water pressure and decreases in effective stress [238]. MICP has shown promising for treatment of liquefiable saturated soils in-situ [239]. MICP can also be used to treat landslide disasters by strengthening sliding surfaces [240]. Furthermore, both deep flooding and heat stress in soil reduced the relative abundance of genes encoding lignin-degrading catalase in Actinobacteria, which resulted in increased organic carbon sequestration [241]. Conversely, drought caused by severe water loss via evaporation can be mitigated with MICP through the formation of a surface crust, remediation of desiccation cracks, smaller pore size, and residual solutes, which decrease the rate of evaporation and water loss by reducing water flow through the material [242].

4. Future Research

While the topic of bacterial carbonatation is not novel, limited research exists on the use of MICP for carbon sequestration. The following is required to better understand microbial carbon sequestration and its long-term feasibility as a CCS technique for the changing environment:
  • Comparisons of MICP utilizing alternative bacteria species to induce different metabolic pathways for the assessment of optimal carbon sequestration.
  • Suitability of specific bacterial species for use with different material types to establish conducive environmental conditions for their metabolic pathways and activity. To date, most MICP research evaluates its usage with soil. However, additional research is required regarding alternative materials that are less hospitable environments for microorganisms to determine the practicality of biochemical carbon sequestration near GHG point-source emissions.
  • Analysis of biochemical alterations for enhancement and optimal use of enzyme-driven metabolic pathways. In addition to optimal growth conditions for bacteria, which is regularly incorporated into biochemical analyses, an evaluation of chemical additives and their impact on the efficacy of metalloenzymes (i.e., Ni and Zn) with the objective of carbon sequestration.
  • Evaluation of MICP and CO2 injection to better understand preferable CO2 phases (liquid, gas, supercritical state) for biocalcification and pressures microorganisms can withstand to maximize the rate-limiting CO2 supply for carbonate precipitation, while minimizing damage to bacterial cells, biomass concentration, and the organic matrix.
  • Comprehensive assessment of bacterial carbonation and its impact on precipitate composition, morphology, and stability for long-term storage of inorganic carbon. Impacts at the micro-particle scale and the large-scale feasibility of carbonate precipitation, considering MICP application and its impact on carbonate stability.
  • Life-cycle assessments of the MICP process comparing different MICP application methods (i.e., in situ biostimulation, ex situ biostimulation, bioaugmentation, amended bioaugmentation) with traditional carbon sequestration techniques to determine quantitively the carbon emissions vs. carbon sequestered from “cradle” to “grave”.
  • Evaluation of the long-term feasibility of MICP with the changing environment due to climate change. The geophysical and biochemical environmental changes (temperature, groundwater conditions, etc.) attributed to climate change and their impacts on specific bacterial species and community diversity, their metabolic activity, and their ability to precipitate carbonates.
MCP and MICP are promising biochemical advancements to the field of carbon sequestration. Research objectives specific to carbon sequestration and CCS are required to further advance the biocalcification methods. Laboratory-scale experimentation and modeling techniques can provide essential information for the development of optimal conditions to maximize carbonate precipitation and the feasibility of combatting climate change.

5. Conclusions

Mineral carbonation is an effective method to store atmospheric CO2 as mineral carbonates via mineral trapping. The mechanism is facilitated by numerous biogenic and abiogenic precipitation techniques with varying degrees of microbial control. Enzyme-driven MICP utilizing CA and urease synergistically is promising for catalysis of CO2 hydration and increase in pH and DIC, leading to increased CaCO3 precipitation. Since increases in CaCO3 precipitation are directly linked to increases in carbon sequestration, optimization of microbial metabolic activity can favorably impact CaCO3 and, therefore, carbon sequestration. Furthermore, solubility trapping and formation trapping via MICP during deep geologic injection of CO2 can mitigate CO2 leakage. Biological carbonate precipitation is promising for atmospheric carbon sequestration with the changing climate. Addition experimental research is required to evaluate the reliability of the method.

Author Contributions

Initial draft, S.M.W.; review and revision of the manuscript, C.N.M. and C.M.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Concordia University and NSERC (grant Number RGPIN-2021-03471).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Calvin, K.; Dasgupta, D.; Krinner, G.; Mukherji, A.; Thorne, P.W.; Trisos, C.; Romero, J.; Aldunce, P.; Barrett, K.; Blanco, G.; et al. IPCC, 2023: Climate Change 2023: Synthesis Report. Contribution of Working Groups I, II and III to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Core Writing Team, Lee, H., Romero, J., Eds.; Intergovernmental Panel on Climate Change (IPCC), 2023; IPCC: Geneva, Switzerland, 2023.
  2. Gayathri, R.; Mahboob, S.; Govindarajan, M.; Al-Ghanim, K.A.; Ahmed, Z.; Al-Mulhm, N.; Vodovnik, M.; Vijayalakshmi, S. A Review on Biological Carbon Sequestration: A Sustainable Solution for a Cleaner Air Environment, Less Pollution and Lower Health Risks. J. King Saud Univ. Sci. 2021, 33, 101282. [Google Scholar] [CrossRef]
  3. Oves, M.; Hussain, F.M.; Ismail, I.M.I.; Felemban, N.M.; Qari, H.A. Microbiological Carbon Sequestration: A Novel Solution for Atmospheric Carbon–Carbon Sequestration through Biological Approach. In Handbook of Research on Inventive Bioremediation Techniques; IGI Global: Hershey, PA, USA, 2017; pp. 108–133. ISBN 978-1-5225-2325-3. [Google Scholar]
  4. Karpiński, P.H.; Bałdyga, J. Chapter 8—Precipitation Processes. In Handbook of Industrial Crystallization; Myerson, A.S., Erdemir, D., Lee, A.Y., Eds.; Cambridge University Press: Cambridge, UK, 2019; pp. 216–265. ISBN 978-1-139-02694-9. [Google Scholar]
  5. Lewis, A. Chapter 4—Precipitation of Heavy Metals. In Sustainable Heavy Metal Remediation: Volume 1: Principles and Processes; Rene, E.R., Sahinkaya, E., Lewis, A., Lens, P.N.L., Eds.; Environmental Chemistry for a Sustainable World; Springer International Publishing: Cham, Switzerland, 2017; pp. 101–120. ISBN 978-3-319-58622-9. [Google Scholar]
  6. The Editors of Encyclopaedia Britannica Chemical Precipitation|Britannica. Available online: https://www.britannica.com/science/chemical-precipitation (accessed on 29 March 2023).
  7. Wang, L.K.; Vaccari, D.A.; Li, Y.; Shammas, N.K. Chapter 5—Chemical Precipitation. In Physicochemical Treatment Processes; Wang, L.K., Hung, Y.-T., Shammas, N.K., Eds.; Handbook of Environmental Engineering; Humana Press: Totowa, NJ, USA, 2005; pp. 141–197. ISBN 978-1-59259-820-5. [Google Scholar]
  8. Yong, R.N.; Mulligan, C.N.; Fukue, M. Sustainable Practices in Geoenvironmental Engineering, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2014; ISBN 978-0-429-16839-0. [Google Scholar]
  9. Chakhmouradian, A.R.; Zaitsev, A.N. Rare Earth Mineralization in Igneous Rocks: Sources and Processes. Elements 2012, 8, 347–353. [Google Scholar] [CrossRef]
  10. Morse, J.W.; Arvidson, R.S. The Dissolution Kinetics of Major Sedimentary Carbonate Minerals. Earth Sci. Rev. 2002, 58, 51–84. [Google Scholar] [CrossRef]
  11. Wada, H.; Tomita, T.; Matsuura, K.; Tuchi, K.; Ito, M.; Morikiyo, T. Graphitization of Carbonaceous Matter during Metamorphism with References to Carbonate and Pelitic Rocks of Contact and Regional Metamorphisms, Japan. Contr. Mineral. Petrol. 1994, 118, 217–228. [Google Scholar] [CrossRef]
  12. James, N.P.; Jones, B. Origin of Carbonate Sedimentary Rocks; John Wiley & Sons: Hoboken, NJ, USA, 2015; ISBN 978-1-118-65270-1. [Google Scholar]
  13. Gregg, J.M.; Bish, D.L.; Kaczmarek, S.E.; Machel, H.G. Mineralogy, Nucleation and Growth of Dolomite in the Laboratory and Sedimentary Environment: A Review. Sedimentology 2015, 62, 1749–1769. [Google Scholar] [CrossRef]
  14. Zhang, G.; Morales, J.; Manuel García-Ruiz, J. Growth Behaviour of Silica/Carbonate Nanocrystalline Composites of Calcite and Aragonite. J. Mater. Chem. B 2017, 5, 1658–1663. [Google Scholar] [CrossRef]
  15. Gonzalez, L.A.; Carpenter, S.J.; Lohmann, K.C. Inorganic Calcite Morphology; Roles of Fluid Chemistry and Fluid Flow. J. Sediment. Res. 1992, 62, 382–399. [Google Scholar] [CrossRef]
  16. Folk, R.L. The Natural History of Crystalline Calcium Carbonate; Effect of Magnesium Content and Salinity. J. Sediment. Res. 1974, 44, 40–53. [Google Scholar] [CrossRef]
  17. Lahann, R.W. A Chemical Model for Calcite Crystal Growth and Morphology Control. J. Sediment. Res. 1978, 48, 337–347. [Google Scholar] [CrossRef]
  18. Zhang, Y.; Dawe, R.A. Influence of Mg2+ on the Kinetics of Calcite Precipitation and Calcite Crystal Morphology. Chem. Geol. 2000, 163, 129–138. [Google Scholar] [CrossRef]
  19. Weyl, P.K. The Change in Solubility of Calcium Carbonate with Temperature and Carbon Dioxide Content. Geochim. Cosmochim. Acta 1959, 17, 214–225. [Google Scholar] [CrossRef]
  20. Brečević, L.; Nielsen, A.E. Solubility of Amorphous Calcium Carbonate. J. Cryst. Growth 1989, 98, 504–510. [Google Scholar] [CrossRef]
  21. Harvey, O.R.; Qafoku, N.P.; Cantrell, K.J.; Lee, G.; Amonette, J.E.; Brown, C.F. Geochemical Implications of Gas Leakage Associated with Geologic CO2 Storage—A Qualitative Review. Environ. Sci. Technol. 2013, 47, 23–36. [Google Scholar] [CrossRef] [PubMed]
  22. Taft, W. Modern Carbonate Sediments. In Carbonate Rocks: Origin, Occurence and Classification; Chilingar, G.V., Bissell, H.J., Fairbridge, R.W., Eds.; Developments in Sedimentology; Elsevier Pub. Co.: Amsterdam, The Netherlands, 1967. [Google Scholar]
  23. Wu, S. Classification of Biogenic Carbonate Rocks. Biopetrology 2021, 1, 19–29. [Google Scholar]
  24. Chilingar, G.V.; Bissell, H.J.; Wolf, K.H. Chapter 5 Diagenesis of Carbonate Rocks. In Developments in Sedimentology; Larsen, G., Chilingar, G.V., Eds.; Diagenesis in Sediments; Elsevier: Amsterdam, The Netherlands, 1967; Volume 8, pp. 179–322. [Google Scholar]
  25. Fairbridge, R.W.; Chilingar, G.V.; Bissell, H.J. Introduction. In Carbonate Rocks: Origin, Occurence and Classification; Chilingarian, G.V., Bissell, H.J., Fairbridge, R.W., Eds.; Developments in Sedimentology; Elsevier Pub. Co.: Amsterdam, The Netherlands, 1967. [Google Scholar]
  26. Castro-Alonso, M.J.; Montañez-Hernandez, L.E.; Sanchez-Muñoz, M.A.; Macias Franco, M.R.; Narayanasamy, R.; Balagurusamy, N. Microbially Induced Calcium Carbonate Precipitation (MICP) and Its Potential in Bioconcrete: Microbiological and Molecular Concepts. Front. Mater. 2019, 6, 126. [Google Scholar] [CrossRef]
  27. Joshi, S.; Goyal, S.; Sudhakara Reddy, M. Influence of Biogenic Treatment in Improving the Durability Properties of Waste Amended Concrete: A Review. Constr. Build. Mater. 2020, 263, 120170. [Google Scholar] [CrossRef]
  28. Kumar, R.; Nongkhlaw, M.; Acharya, C.; Joshi, S.R. Bacterial Community Structure from the Perspective of the Uranium Ore Deposits of Domiasiat in India. Proc. Natl. Acad. Sci. India Sect. B Biol. Sci. 2013, 83, 485–497. [Google Scholar] [CrossRef]
  29. Weiner, S.; Dove, P.M. An Overview of Biomineralization Processes and the Problem of the Vital Effect. Rev. Mineral. Geochem. 2003, 54, 1–29. [Google Scholar] [CrossRef]
  30. Zhuang, D.; Yan, H.; Tucker, M.E.; Zhao, H.; Han, Z.; Zhao, Y.; Sun, B.; Li, D.; Pan, J.; Zhao, Y.; et al. Calcite Precipitation Induced by Bacillus Cereus MRR2 Cultured at Different Ca2+ Concentrations: Further Insights into Biotic and Abiotic Calcite. Chem. Geol. 2018, 500, 64–87. [Google Scholar] [CrossRef]
  31. Kelts, K.; Talbot, M. Lacustrine Carbonates as Geochemical Archives of Environmental Change and Biotic/Abiotic Interactions. In Large Lakes: Ecological Structure and Function; Tilzer, M.M., Serruya, C., Eds.; Springer: Berlin/Heidelberg, Germany, 1990; pp. 288–315. ISBN 978-3-642-84077-7. [Google Scholar]
  32. Rainey, D.K.; Jones, B. Abiotic versus Biotic Controls on the Development of the Fairmont Hot Springs Carbonate Deposit, British Columbia, Canada. Sedimentology 2009, 56, 1832–1857. [Google Scholar] [CrossRef]
  33. Okyay, T.O.; Rodrigues, D.F. Biotic and Abiotic Effects on CO2 Sequestration during Microbially-Induced Calcium Carbonate Precipitation. FEMS Microbiol. Ecol. 2015, 91, fiv017. [Google Scholar] [CrossRef] [PubMed]
  34. Anbu, P.; Kang, C.-H.; Shin, Y.-J.; So, J.-S. Formations of Calcium Carbonate Minerals by Bacteria and Its Multiple Applications. Springerplus 2016, 5, 250. [Google Scholar] [CrossRef] [PubMed]
  35. Mujah, D.; Shahin, M.; Cheng, L. State-of-the-Art Review of Biocementation by Microbially Induced Calcite Precipitation (MICP) for Soil Stabilization. Geomicrobiology 2016, 34, 524–537. [Google Scholar] [CrossRef]
  36. Dhami, N.K.; Reddy, M.S.; Mukherjee, A. Biomineralization of Calcium Carbonates and Their Engineered Applications: A Review. Front. Microbiol. 2013, 4, 314. [Google Scholar] [CrossRef]
  37. Benzerara, K.; Miot, J.; Morin, G.; Ona-Nguema, G.; Skouri-Panet, F.; Férard, C. Significance, Mechanisms and Environmental Implications of Microbial Biomineralization. Comptes Rendus Geosci. 2011, 343, 160–167. [Google Scholar] [CrossRef]
  38. Dupraz, C.; Reid, R.P.; Braissant, O.; Decho, A.W.; Norman, R.S.; Visscher, P.T. Processes of Carbonate Precipitation in Modern Microbial Mats. Earth Sci. Rev. 2009, 96, 141–162. [Google Scholar] [CrossRef]
  39. De Deyn, G.B.; Cornelissen, J.H.C.; Bardgett, R.D. Plant Functional Traits and Soil Carbon Sequestration in Contrasting Biomes. Ecol. Lett. 2008, 11, 516–531. [Google Scholar] [CrossRef]
  40. Lal, R. Carbon Sequestration. Philos. Trans. R. Soc. B Biol. Sci. 2007, 363, 815–830. [Google Scholar] [CrossRef]
  41. Farrelly, D.J.; Everard, C.D.; Fagan, C.C.; McDonnell, K.P. Carbon Sequestration and the Role of Biological Carbon Mitigation: A Review. Renew. Sustain. Energy Rev. 2013, 21, 712–727. [Google Scholar] [CrossRef]
  42. Liu, D.-X.; Mai, Z.-M.; Sun, C.-C.; Zhou, Y.-W.; Liao, H.-H.; Wang, Y.-S.; Cheng, H. Dynamics of Extracellular Polymeric Substances and Soil Organic Carbon with Mangrove Zonation along a Continuous Tidal Gradient. Front. Mar. Sci. 2022, 9, 967767. [Google Scholar] [CrossRef]
  43. Alongi, D.M. Carbon Cycling and Storage in Mangrove Forests. Annu. Rev. Mar. Sci. 2014, 6, 195–219. [Google Scholar] [CrossRef] [PubMed]
  44. Konhauser, K. Introduction to Geomicrobiology; Blackwell Publishing: Oxford, UK, 2007; ISBN 978-0-632-05454-1. [Google Scholar]
  45. Wilcox, S.M.; Mulligan, C.N.; Neculita, C.M. Microbially Induced Calcium Carbonate Precipitation as a Bioremediation Technique for Mining Waste. Toxics 2024, 12, 107. [Google Scholar] [CrossRef] [PubMed]
  46. Mitchell, A.C.; Phillips, A.J.; Hiebert, R.; Gerlach, R.; Spangler, L.H.; Cunningham, A.B. Biofilm Enhanced Geologic Sequestration of Supercritical CO2. Int. J. Greenh. Gas Control 2009, 3, 90–99. [Google Scholar] [CrossRef]
  47. Jiang, N.-J.; Wang, Y.-J.; Chu, J.; Kawasaki, S.; Tang, C.-S.; Cheng, L.; Du, Y.-J.; Shashank, B.S.; Singh, D.N.; Han, X.-L.; et al. Bio-Mediated Soil Improvement: An Introspection into Processes, Materials, Characterization and Applications. Soil Use Manag. 2022, 38, 68–93. [Google Scholar] [CrossRef]
  48. Madigan, M.T.; Bender, K.S.; Buckley, D.H.; Sattley, W.M.; Stahl, D.A. Brock Biology of Microorganisms, 15th ed.; Global Edition; Pearson: New York, NY, USA, 2019; ISBN 978-1-292-23510-3. [Google Scholar]
  49. Westbroek, P.; Brown, C.W.; van Bleijswijk, J.; Brownlee, C.; Brummer, G.J.; Conte, M.; Egge, J.; Fernández, E.; Jordan, R.; Knappertsbusch, M.; et al. A Model System Approach to Biological Climate Forcing. The Example of Emiliania Huxleyi. Glob. Planet. Change 1993, 8, 27–46. [Google Scholar] [CrossRef]
  50. Riebesell, U.; Zondervan, I.; Rost, B.; Tortell, P.D.; Zeebe, R.E.; Morel, F.M.M. Reduced Calcification of Marine Plankton in Response to Increased Atmospheric CO2. Nature 2000, 407, 364–367. [Google Scholar] [CrossRef]
  51. Berg, I.A.; Kockelkorn, D.; Ramos-Vera, W.H.; Say, R.F.; Zarzycki, J.; Hügler, M.; Alber, B.E.; Fuchs, G. Autotrophic Carbon Fixation in Archaea. Nat. Rev. Microbiol 2010, 8, 447–460. [Google Scholar] [CrossRef]
  52. Hammes, F.; Verstraete, W. Key Roles of pH and Calcium Metabolism in Microbial Carbonate Precipitation. Rev. Environ. Sci. Biotechnol. 2002, 1, 3–7. [Google Scholar] [CrossRef]
  53. Phillips, A.J.; Gerlach, R.; Lauchnor, E.; Mitchell, A.C.; Cunningham, A.B.; Spangler, L. Engineered Applications of Ureolytic Biomineralization: A Review. Biofouling 2013, 29, 715–733. [Google Scholar] [CrossRef]
  54. Achal, V.; Pan, X. Characterization of Urease and Carbonic Anhydrase Producing Bacteria and Their Role in Calcite Precipitation. Curr. Microbiol. 2011, 62, 894–902. [Google Scholar] [CrossRef]
  55. Douglas, S.; Beveridge, T.J. Mineral Formation by Bacteria in Natural Microbial Communities. FEMS Microbiol. Ecol. 1998, 26, 79–88. [Google Scholar] [CrossRef]
  56. Kapahi, M.; Sachdeva, S. Bioremediation Options for Heavy Metal Pollution. J. Health Pollut. 2019, 9, 191203. [Google Scholar] [CrossRef] [PubMed]
  57. Yee, N.; Fein, J.B.; Daughney, C.J. Experimental Study of the pH, Ionic Strength, and Reversibility Behavior of Bacteria–Mineral Adsorption. Geochim. Cosmochim. Acta 2000, 64, 609–617. [Google Scholar] [CrossRef]
  58. Zhu, T.; Dittrich, M. Carbonate Precipitation through Microbial Activities in Natural Environment, and Their Potential in Biotechnology: A Review. Front. Bioeng. Biotechnol. 2016, 4, 4. [Google Scholar] [CrossRef]
  59. Xiangliang, P. Micrologically Induced Carbonate Precipitation as a Promising Way to in Situ Immobilize Heavy Metals in Groundwater and Sediment. Res. J. Chem. Environ. 2009, 13, 3–4. [Google Scholar]
  60. Achal, V.; Mukherjee, A.; Basu, P.C.; Reddy, M.S. Strain Improvement of Sporosarcina pasteurii for Enhanced Urease and Calcite Production. J. Ind. Microbiol. Biotechnol. 2009, 36, 981–988. [Google Scholar] [CrossRef]
  61. Ferris, F.G.; Fyfe, W.S.; Beveridge, T.J. Bacteria as Nucleation Sites for Authigenic Minerals in a Metal-Contaminated Lake Sediment. Chem. Geol. 1987, 63, 225–232. [Google Scholar] [CrossRef]
  62. Costa, O.Y.A.; Raaijmakers, J.M.; Kuramae, E.E. Microbial Extracellular Polymeric Substances: Ecological Function and Impact on Soil Aggregation. Front. Microbiol. 2018, 9, 1636. [Google Scholar] [CrossRef]
  63. Dittrich, M.; Sibler, S. Calcium Carbonate Precipitation by Cyanobacterial Polysaccharides. Geol. Soc. Lond. Spec. Publ. 2010, 336, 51–63. [Google Scholar] [CrossRef]
  64. Tourney, J.; Ngwenya, B.T.; Mosselmans, J.W.F.; Tetley, L.; Cowie, G.L. The Effect of Extracellular Polymers (EPS) on the Proton Adsorption Characteristics of the Thermophile Bacillus Licheniformis S-86. Chem. Geol. 2008, 247, 1–15. [Google Scholar] [CrossRef]
  65. Kawaguchi, T.; Decho, A.W. A Laboratory Investigation of Cyanobacterial Extracellular Polymeric Secretions (EPS) in Influencing CaCO3 Polymorphism. J. Cryst. Growth 2002, 240, 230–235. [Google Scholar] [CrossRef]
  66. Kim, J.-H.; Lee, J.-Y. An Optimum Condition of MICP Indigenous Bacteria with Contaminated Wastes of Heavy Metal. J. Mater. Cycles Waste Manag. 2019, 21, 239–247. [Google Scholar] [CrossRef]
  67. Tourney, J.; Ngwenya, B.T. Bacterial Extracellular Polymeric Substances (EPS) Mediate CaCO3 Morphology and Polymorphism. Chem. Geol. 2009, 262, 138–146. [Google Scholar] [CrossRef]
  68. Dejong, J.t.; Soga, K.; Kavazanjian, E.; Burns, S.; Van Paassen, L.a.; Al Qabany, A.; Aydilek, A.; Bang, S.s.; Burbank, M.; Caslake, L.f.; et al. Biogeochemical Processes and Geotechnical Applications: Progress, Opportunities and Challenges. Géotechnique 2013, 63, 287–301. [Google Scholar] [CrossRef]
  69. Achal, V.; Li, M.; Zhang, Q. Biocement, Recent Research in Construction Engineering: Status of China against Rest of World. Adv. Cem. Res. 2014, 26, 281–291. [Google Scholar] [CrossRef]
  70. Fenchel, T.; King, G.M.; Blackburn, T.H. Bacterial Biogeochemistry: The Ecophysiology of Mineral Cycling; Academic Press: Cambridge, MA, USA, 2012; ISBN 978-0-12-415836-8. [Google Scholar]
  71. Fontaine, S.; Mariotti, A.; Abbadie, L. The Priming Effect of Organic Matter: A Question of Microbial Competition? Soil Biol. Biochem. 2003, 35, 837–843. [Google Scholar] [CrossRef]
  72. Pepper, I.L.; Gerba, C.P.; Gentry, T.J. Environmental Microbiology, 3rd ed.; Elsevier: Amsterdam, The Netherlands, 2015; ISBN 978-0-12-394626-3. [Google Scholar]
  73. Govarthanan, M.; Lee, K.-J.; Cho, M.; Kim, J.S.; Kamala-Kannan, S.; Oh, B.-T. Significance of Autochthonous bacillus sp. KK1 on Biomineralization of Lead in Mine Tailings. Chemosphere 2013, 90, 2267–2272. [Google Scholar] [CrossRef]
  74. Bang, S. Microbiologically-Enhanced Crack Remediation (MECR). Proc. Microbiol. Soc. Korea Conf. 2001, 11, 26–36. [Google Scholar]
  75. Chuo, S.C.; Mohamed, S.F.; Mohd Setapar, S.H.; Ahmad, A.; Jawaid, M.; Wani, W.A.; Yaqoob, A.A.; Mohamad Ibrahim, M.N. Insights into the Current Trends in the Utilization of Bacteria for Microbially Induced Calcium Carbonate Precipitation. Materials 2020, 13, 4993. [Google Scholar] [CrossRef]
  76. Graddy, C.M.R.; Gomez, M.G.; DeJong, J.T.; Nelson, D.C. Native Bacterial Community Convergence in Augmented and Stimulated Ureolytic MICP Biocementation. Environ. Sci. Technol. 2021, 55, 10784–10793. [Google Scholar] [CrossRef]
  77. McConnaughey, T.A.; Whelan, J.F. Calcification Generates Protons for Nutrient and Bicarbonate Uptake. Earth Sci. Rev. 1997, 42, 95–117. [Google Scholar] [CrossRef]
  78. Castanier, S.; Le Métayer-Levrel, G.; Perthuisot, J.-P. Ca-Carbonates Precipitation and Limestone Genesis—The Microbiogeologist Point of View. Sediment. Geol. 1999, 126, 9–23. [Google Scholar] [CrossRef]
  79. Norris, V.; Grant, S.; Freestone, P.; Canvin, J.; Sheikh, F.N.; Toth, I.; Trinei, M.; Modha, K.; Norman, R.I. Calcium Signalling in Bacteria. J. Bacteriol. 1996, 178, 3677–3682. [Google Scholar] [CrossRef]
  80. Kumar, A.; Song, H.-W.; Mishra, S.; Zhang, W.; Zhang, Y.-L.; Zhang, Q.-R.; Yu, Z.-G. Application of Microbial-Induced Carbonate Precipitation (MICP) Techniques to Remove Heavy Metal in the Natural Environment: A Critical Review. Chemosphere 2023, 318, 137894. [Google Scholar] [CrossRef]
  81. Rahman, M.M.; Hora, R.N.; Ahenkorah, I.; Beecham, S.; Karim, M.R.; Iqbal, A. State-of-the-Art Review of Microbial-Induced Calcite Precipitation and Its Sustainability in Engineering Applications. Sustainability 2020, 12, 6281. [Google Scholar] [CrossRef]
  82. Olaya-Abril, A.; Hidalgo-Carrillo, J.; Luque-Almagro, V.M.; Fuentes-Almagro, C.; Urbano, F.J.; Moreno-Vivián, C.; Richardson, D.J.; Roldán, M.D. Exploring the Denitrification Proteome of Paracoccus Denitrificans PD1222. Front. Microbiol. 2018, 9, 1137. [Google Scholar] [CrossRef]
  83. van Paassen, L.A.; Daza, C.M.; Staal, M.; Sorokin, D.Y.; van der Zon, W.; van Loosdrecht, M.C.M. Potential Soil Reinforcement by Biological Denitrification. Ecol. Eng. 2010, 36, 168–175. [Google Scholar] [CrossRef]
  84. Hamdan, N.; Kavazanjian, E.; Rittmann, B.E.; Karatas, I. Carbonate Mineral Precipitation for Soil Improvement through Microbial Denitrification. Geomicrobiol. J. 2017, 34, 139–146. [Google Scholar] [CrossRef]
  85. Liang, Y.; Fu, R.; Sailike, A.; Hao, H.; Yu, Z.; Wang, R.; Peng, N.; Li, S.; Zhang, W.; Liu, Y. Soil Labile Organic Carbon and Nitrate Nitrogen Are the Main Factors Driving Carbon-Fixing Pathways during Vegetation Restoration in the Loess Plateau, China. Agric. Ecosyst. Environ. 2025, 378, 109283. [Google Scholar] [CrossRef]
  86. Ren, M.; Zhang, Z.; Wang, X.; Zhou, Z.; Chen, D.; Zeng, H.; Zhao, S.; Chen, L.; Hu, Y.; Zhang, C.; et al. Diversity and Contributions to Nitrogen Cycling and Carbon Fixation of Soil Salinity Shaped Microbial Communities in Tarim Basin. Front. Microbiol. 2018, 9, 431. [Google Scholar] [CrossRef]
  87. Chekroun, K.B.; Rodríguez-Navarro, C.; González-Muñoz, M.T.; Arias, J.M.; Cultrone, G.; Rodríguez-Gallego, M. Precipitation and Growth Morphology of Calcium Carbonate Induced by Myxococcus xanthus: Implications for Recognition of Bacterial Carbonates. J. Sediment. Res. 2004, 74, 868–876. [Google Scholar] [CrossRef]
  88. Jiménez-López, C.; Jroundi, F.; Rodríguez-Gallego, M.; Arias, J.; Gonzalez, M.T. Biomineralization Induced by Myxobacteria. In Communicating Current Research and Educational Topics and Trends in Applied Microbiology; Méndez-Vilas, A., Ed.; Formatex Research Center: Oviedo, Spain, 2007. [Google Scholar]
  89. Marín-Ortega, S.; Torras, M.À.C.i.; Iglesias-Campos, M.Á. Microbially Induced Calcium Carbonate Precipitation in Fossil Consolidation Treatments: Preliminary Results Inducing Exogenous Myxococcus xanthus Bacteria in a Miocene Cheirogaster Richardi Specimen. Heliyon 2023, 9, e17597. [Google Scholar] [CrossRef] [PubMed]
  90. Keena, M.; Meehan, M.; Scherer, T. Nitrogen Behaviour in the Environment; North Dakota State University: Fargo, ND, USA, 2022. [Google Scholar]
  91. Lee, M.; Gomez, M.G.; San Pablo, A.C.M.; Kolbus, C.M.; Graddy, C.M.R.; DeJong, J.T.; Nelson, D.C. Investigating Ammonium by-Product Removal for Ureolytic Bio-Cementation Using Meter-Scale Experiments. Sci. Rep. 2019, 9, 18313. [Google Scholar] [CrossRef] [PubMed]
  92. Joshi, S.; Goyal, S.; Mukherjee, A.; Reddy, M.S. Microbial Healing of Cracks in Concrete: A Review. J. Ind. Microbiol. Biotechnol. 2017, 44, 1511–1525. [Google Scholar] [CrossRef]
  93. Dhami, N.K.; Reddy, M.S.; Mukherjee, A. Synergistic Role of Bacterial Urease and Carbonic Anhydrase in Carbonate Mineralization. Appl. Biochem. Biotechnol. 2014, 172, 2552–2561. [Google Scholar] [CrossRef]
  94. Stocks-Fischer, S.; Galinat, J.K.; Bang, S.S. Microbiological Precipitation of CaCO3. Soil Biol. Biochem. 1999, 31, 1563–1571. [Google Scholar] [CrossRef]
  95. Hasan, H.A.H. Ureolytic Microorganisms and Soil Fertility: A Review. Commun. Soil Sci. Plant Anal. 2000, 31, 2565–2589. [Google Scholar] [CrossRef]
  96. Zerner, B. Recent Advances in the Chemistry of an Old Enzyme, Urease. Bioorganic Chem. 1991, 19, 116–131. [Google Scholar] [CrossRef]
  97. Callahan, B.P.; Yuan, Y.; Wolfenden, R. The Burden Borne by Urease. J. Am. Chem. Soc. 2005, 127, 10828–10829. [Google Scholar] [CrossRef]
  98. Ciurli, S. Urease: Recent Insights on the Role of Nickel. In Nickel and Its Surprising Impact in Nature; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2007; pp. 241–277. ISBN 978-0-470-02813-1. [Google Scholar]
  99. Su, F.; Yang, Y.Y. Microbially Induced Carbonate Precipitation via Methanogenesis Pathway by a Microbial Consortium Enriched from Activated Anaerobic Sludge. J. Appl. Microbiol. 2021, 131, 236–256. [Google Scholar] [CrossRef]
  100. DeJong, J.; Proto, C.; Kuo, M.; Gomez, M. Bacteria, Biofilms, and Invertebrates: The next Generation of Geotechnical Engineers? In Proceedings of the Geo-Congress 2014, Atlanta, GA, USA, 23–26 February 2014; pp. 3959–3968. [Google Scholar] [CrossRef]
  101. Buikema, N.D.; Zwissler, B.E.; Seagren, E.A.; Oommen, T.; Vitton, S. Stabilisation of Iron Mine Tailings through Biocalcification. Environ. Geotech. 2018, 5, 94–106. [Google Scholar] [CrossRef]
  102. Stabnikov, V.; Ivanov, V. Biotechnological Production of Biogrout from Iron Ore and Cellulose. J. Chem. Technol. Biotechnol. 2017, 92, 180–187. [Google Scholar] [CrossRef]
  103. Yao, Y.; Wang, L.; Hemamali Peduruhewa, J.; Van Zwieten, L.; Gong, L.; Tan, B.; Zhang, G. The Coupling between Iron and Carbon and Iron Reducing Bacteria Control Carbon Sequestration in Paddy Soils. CATENA 2023, 223, 106937. [Google Scholar] [CrossRef]
  104. Chen, C.; Dynes, J.J.; Wang, J.; Sparks, D.L. Properties of Fe-Organic Matter Associations via Coprecipitation versus Adsorption. Environ. Sci. Technol. 2014, 48, 13751–13759. [Google Scholar] [CrossRef] [PubMed]
  105. Lalonde, K.; Mucci, A.; Ouellet, A.; Gélinas, Y. Preservation of Organic Matter in Sediments Promoted by Iron. Nature 2012, 483, 198–200. [Google Scholar] [CrossRef]
  106. Bergdale, T.E.; Pinkelman, R.J.; Hughes, S.R.; Zambelli, B.; Ciurli, S.; Bang, S.S. Engineered Biosealant Strains Producing Inorganic and Organic Biopolymers. J. Biotechnol. 2012, 161, 181–189. [Google Scholar] [CrossRef]
  107. Zambelli, B.; Musiani, F.; Benini, S.; Ciurli, S. Chemistry of Ni2+ in Urease: Sensing, Trafficking, and Catalysis. Acc. Chem. Res. 2011, 44, 520–530. [Google Scholar] [CrossRef]
  108. Svane, S.; Sigurdarson, J.J.; Finkenwirth, F.; Eitinger, T.; Karring, H. Inhibition of Urease Activity by Different Compounds Provides Insight into the Modulation and Association of Bacterial Nickel Import and Ureolysis. Sci. Rep. 2020, 10, 8503. [Google Scholar] [CrossRef]
  109. Lv, J.; Jiang, Y.; Yu, Q.; Lu, S. Structural and Functional Role of Nickel Ions in Urease by Molecular Dynamics Simulation. J. Biol. Inorg. Chem. 2011, 16, 125–135. [Google Scholar] [CrossRef]
  110. Estiu, G.; Merz, K.M. Catalyzed Decomposition of Urea. Molecular Dynamics Simulations of the Binding of Urea to Urease. Biochemistry 2006, 45, 4429–4443. [Google Scholar] [CrossRef]
  111. Bachmeier, K.L.; Williams, A.E.; Warmington, J.R.; Bang, S.S. Urease Activity in Microbiologically-Induced Calcite Precipitation. J. Biotechnol. 2002, 93, 171–181. [Google Scholar] [CrossRef] [PubMed]
  112. Okyay, T.O.; Rodrigues, D.F. Optimized Carbonate Micro-Particle Production by Sporosarcina pasteurii Using Response Surface Methodology. Ecol. Eng. 2014, 62, 168–174. [Google Scholar] [CrossRef]
  113. Okyay, T.O.; Nguyen, H.N.; Castro, S.L.; Rodrigues, D.F. CO2 Sequestration by Ureolytic Microbial Consortia through Microbially-Induced Calcite Precipitation. Sci. Total Environ. 2016, 572, 671–680. [Google Scholar] [CrossRef]
  114. Kaur, G.; Dhami, N.K.; Goyal, S.; Mukherjee, A.; Reddy, M.S. Utilization of Carbon Dioxide as an Alternative to Urea in Biocementation. Constr. Build. Mater. 2016, 123, 527–533. [Google Scholar] [CrossRef]
  115. Sánchez-Andrea, I.; Sanz, J.L.; Bijmans, M.F.M.; Stams, A.J.M. Sulfate Reduction at Low pH to Remediate Acid Mine Drainage. J. Hazard. Mater. 2014, 269, 98–109. [Google Scholar] [CrossRef]
  116. Sánchez-Andrea, I.; Stams, A.J.M.; Weijma, J.; Gonzalez Contreras, P.; Dijkman, H.; Rozendal, R.A.; Johnson, D.B. A Case in Support of Implementing Innovative Bio-Processes in the Metal Mining Industry. FEMS Microbiol. Lett. 2016, 363, 1–4. [Google Scholar] [CrossRef]
  117. Paul, V.G.; Wronkiewicz, D.J.; Mormile, M.R. Impact of Elevated CO2 Concentrations on Carbonate Mineral Precipitation Ability of Sulfate-Reducing Bacteria and Implications for CO2 Sequestration. Appl. Geochem. 2017, 78, 250–271. [Google Scholar] [CrossRef]
  118. Pester, M.; Knorr, K.-H.; Friedrich, M.W.; Wagner, M.; Loy, A. Sulfate-Reducing Microorganisms in Wetlands—Fameless Actors in Carbon Cycling and Climate Change. Front. Microbiol. 2012, 3, 19769. [Google Scholar] [CrossRef]
  119. Visscher, P.T.; Reid, R.P.; Bebout, B.M.; Hoeft, S.E.; Macintyre, I.G.; Thompson, J.A. Formation of Lithified Micritic Laminae in Modern Marine Stromatolites (Bahamas); The Role of Sulfur Cycling. Am. Mineral. 1998, 83, 1482–1493. [Google Scholar] [CrossRef]
  120. Görgen, S.; Benzerara, K.; Skouri-Panet, F.; Gugger, M.; Chauvat, F.; Cassier-Chauvat, C. The Diversity of Molecular Mechanisms of Carbonate Biomineralization by Bacteria. Discov. Mater. 2020, 1, 2. [Google Scholar] [CrossRef]
  121. Akam, S.A.; Swanner, E.D.; Yao, H.; Hong, W.-L.; Peckmann, J. Methane-Derived Authigenic Carbonates—A Case for a Globally Relevant Marine Carbonate Factory. Earth Sci. Rev. 2023, 243, 104487. [Google Scholar] [CrossRef]
  122. Caesar, K.H.; Kyle, J.R.; Lyons, T.W.; Tripati, A.; Loyd, S.J. Carbonate Formation in Salt Dome Cap Rocks by Microbial Anaerobic Oxidation of Methane. Nat. Commun. 2019, 10, 808. [Google Scholar] [CrossRef] [PubMed]
  123. Jain, S.; Fang, C.; Achal, V. A Critical Review on Microbial Carbonate Precipitation via Denitrification Process in Building Materials. Bioengineered 2021, 12, 7529–7551. [Google Scholar] [CrossRef] [PubMed]
  124. Rasigraf, O.; Kool, D.M.; Jetten, M.S.M.; Sinninghe Damsté, J.S.; Ettwig, K.F. Autotrophic Carbon Dioxide Fixation via the Calvin-Benson-Bassham Cycle by the Denitrifying Methanotroph “Candidatus Methylomirabilis Oxyfera”. Appl. Environ. Microbiol. 2014, 80, 2451–2460. [Google Scholar] [CrossRef]
  125. Smith, K.S.; Ferry, J.G. Prokaryotic Carbonic Anhydrases. FEMS Microbiol. Rev. 2000, 24, 335–366. [Google Scholar] [CrossRef]
  126. Altermann, W.; Kazmierczak, J.; Oren, A.; Wright, D.T. Cyanobacterial Calcification and Its Rock-Building Potential during 3.5 Billion Years of Earth History. Geobiology 2006, 4, 147–166. [Google Scholar] [CrossRef]
  127. Verrecchia, E.P.; Freytet, P.; Verrecchia, K.E.; Dumont, J.-L. Spherulites in Calcrete Laminar Crusts; Biogenic CaCO 3 Precipitation as a Major Contributor to Crust Formation. J. Sediment. Res. 1995, 65, 690–700. [Google Scholar] [CrossRef]
  128. Pentecost, A. The Formation of Travertine Shrubs: Mammoth Hot Springs, Wyoming. Geol. Mag. 1990, 127, 159–168. [Google Scholar] [CrossRef]
  129. Jones, T.R.; Poitras, J.; Levett, A.; Langendam, A.; Vietti, A.; Southam, G. Accelerated Carbonate Biomineralisation of Venetia Diamond Mine Coarse Residue Deposit (CRD) Material—A Field Trial Study. Sci. Total Environ. 2023, 893, 164853. [Google Scholar] [CrossRef]
  130. Jones, T.R.; Poitras, J.; Paterson, D.; Southam, G. Historical Diamond Mine Waste Reveals Carbon Sequestration Resource in Kimberlite Residue. Chem. Geol. 2023, 617, 121270. [Google Scholar] [CrossRef]
  131. Sundaram, S.; Thakur, I.S. Induction of Calcite Precipitation through Heightened Production of Extracellular Carbonic Anhydrase by CO2 Sequestering Bacteria. Bioresour. Technol. 2018, 253, 368–371. [Google Scholar] [CrossRef] [PubMed]
  132. Lam, L.; Ilies, M.A. Evaluation of the Impact of Esterases and Lipases from the Circulatory System against Substrates of Different Lipophilicity. Int. J. Mol. Sci. 2022, 23, 1262. [Google Scholar] [CrossRef] [PubMed]
  133. Kaplan, A.; Reinhold, L. CO2 Concentrating Mechanisms in Photosynthetic Microorganisms. Annu. Rev. Plant Biol. 1999, 50, 539–570. [Google Scholar] [CrossRef] [PubMed]
  134. Supuran, C.T.; Capasso, C. An Overview of the Bacterial Carbonic Anhydrases. Metabolites 2017, 7, 56. [Google Scholar] [CrossRef]
  135. Badger, M.R.; Hanson, D.; Price, G.D. Evolution and Diversity of CO2 Concentrating Mechanisms in Cyanobacteria. Funct. Plant Biol. 2002, 29, 161–173. [Google Scholar] [CrossRef]
  136. Klanchui, A.; Cheevadhanarak, S.; Prommeenate, P.; Meechai, A. Exploring Components of the CO2-Concentrating Mechanism in Alkaliphilic Cyanobacteria Through Genome-Based Analysis. Comput. Struct. Biotechnol. J. 2017, 15, 340–350. [Google Scholar] [CrossRef]
  137. Price, G.D.; Badger, M.R.; Woodger, F.J.; Long, B.M. Advances in Understanding the Cyanobacterial CO2-Concentrating-Mechanism (CCM): Functional Components, Ci Transporters, Diversity, Genetic Regulation and Prospects for Engineering into Plants. J. Exp. Bot. 2008, 59, 1441–1461. [Google Scholar] [CrossRef]
  138. Badger, M.R.; Price, G.D.; Long, B.M.; Woodger, F.J. The Environmental Plasticity and Ecological Genomics of the Cyanobacterial CO2 Concentrating Mechanism. J. Exp. Bot. 2006, 57, 249–265. [Google Scholar] [CrossRef]
  139. Capasso, C.; Supuran, C.T. An Overview of the Alpha-, Beta- and Gamma-Carbonic Anhydrases from Bacteria: Can Bacterial Carbonic Anhydrases Shed New Light on Evolution of Bacteria? J. Enzym. Inhib. Med. Chem. 2015, 30, 325–332. [Google Scholar] [CrossRef]
  140. Sharma, A.; Bhattacharya, A.; Singh, S. Purification and Characterization of an Extracellular Carbonic Anhydrase from Pseudomonas Fragi. Process Biochem. 2009, 44, 1293–1297. [Google Scholar] [CrossRef]
  141. Kupriyanova, E.; Villarejo, A.; Markelova, A.; Gerasimenko, L.; Zavarzin, G.; Samuelsson, G.; Los, D.A.; Pronina, N. Extracellular Carbonic Anhydrases of the Stromatolite-Forming Cyanobacterium Microcoleus Chthonoplastes. Microbiology 2007, 153, 1149–1156. [Google Scholar] [CrossRef] [PubMed]
  142. Zhang, Z.; Lian, B.; Chen, M.; Li, X.; Li, Y. Bacillus Mucilaginosus Can Capture Atmospheric CO2 by Carbonic Anhydrase. Afr. J. Microbiol. Res. 2011, 5, 106–112. [Google Scholar]
  143. Li, W.; Yu, L.; He, Q.; Wu, Y.; Yuan, D.; Cao, J. Effects of Microbes and Their Carbonic Anhydrase on Ca2+ and Mg2+ Migration in Column-Built Leached Soil-Limestone Karst Systems. Appl. Soil Ecol. 2005, 29, 274–281. [Google Scholar] [CrossRef]
  144. Nathan, V.K.; Ammini, P. Carbon Dioxide Sequestering Ability of Bacterial Carbonic Anhydrase in a Mangrove Soil Microcosm and Its Bio-Mineralization Properties. Water Air Soil Pollut. 2019, 230, 192. [Google Scholar] [CrossRef]
  145. Zhu, J.; Sun, J.; Pang, C.; Li, Q.; Yang, Z.; Li, G. Isolation, Identification, and Carbonate Mineralization Characteristics of a Newly Carbonic Anhydrase-Producing Strain. Appl. Biochem. Biotechnol. 2024, 196, 8009–8025. [Google Scholar] [CrossRef]
  146. Christianson, D.W.; Fierke, C.A. Carbonic Anhydrase:  Evolution of the Zinc Binding Site by Nature and by Design. Acc. Chem. Res. 1996, 29, 331–339. [Google Scholar] [CrossRef]
  147. Kiefer, L.L.; Paterno, S.A.; Fierke, C.A. Hydrogen Bond Network in the Metal Binding Site of Carbonic Anhydrase Enhances Zinc Affinity and Catalytic Efficiency. J. Am. Chem. Soc. 1995, 117, 6831–6837. [Google Scholar] [CrossRef]
  148. Supuran, C.T.; Scozzafava, A.; Casini, A. Carbonic Anhydrase Inhibitors. Med. Res. Rev. 2003, 23, 146–189. [Google Scholar] [CrossRef]
  149. Tupper, R.; Watts, R.W.E.; Wormall, A. Some Observations on the Zinc in Carbonic Anhydrase. Biochem. J. 1952, 50, 429–432. [Google Scholar] [CrossRef]
  150. Håkansson, K.; Carlsson, M.; Svensson, L.A.; Liljas, A. Structure of Native and Apo Carbonic Anhydrase II and Structure of Some of Its Anion-Ligand Complexes. J. Mol. Biol. 1992, 227, 1192–1204. [Google Scholar] [CrossRef]
  151. McCall, K.A.; Huang, C.; Fierke, C.A. Function and Mechanism of Zinc Metalloenzymes. J. Nutr. 2000, 130, 1437S–1446S. [Google Scholar] [CrossRef] [PubMed]
  152. Dhami, N.; Reddy, M.; Mukherjee, A. Biomineralization of Calcium Carbonate Polymorphs by the Bacterial Strains Isolated from Calcareous Sites. J. Microbiol. Biotechnol. 2013, 23, 707–714. [Google Scholar] [CrossRef] [PubMed]
  153. Zhao, J.; Lu, W.; Zhang, F.; Lu, C.; Du, J.; Zhu, R.; Sun, L. Evaluation of CO2 Solubility-Trapping and Mineral-Trapping in Microbial-Mediated CO2–Brine–Sandstone Interaction. Mar. Pollut. Bull. 2014, 85, 78–85. [Google Scholar] [CrossRef]
  154. Achal, V.; Pan, X.; Lee, D.-J.; Kumari, D.; Zhang, D. Remediation of Cr(VI) from Chromium Slag by Biocementation. Chemosphere 2013, 93, 1352–1358. [Google Scholar] [CrossRef]
  155. Park, I.-S.; Hausinger, R.P. Requirement of Carbon Dioxide for in Vitro Assembly of the Urease Nickel Metallocenter. Science 1995, 267, 1156–1158. [Google Scholar] [CrossRef]
  156. Lal, R. Forest Soils and Carbon Sequestration. For. Ecol. Manag. 2005, 220, 242–258. [Google Scholar] [CrossRef]
  157. Dhanwantri, K.; Sharma, P.; Mehta, S.; Prakash, P. Carbon Sequestration, Its Methods and Significance. In Environmental Sustainability: Concepts, Principles, Evidences and Innovations; Mishra, G.C., Ed.; Excellent Publishing House: New Delhi, India, 2014; pp. 94–98. ISBN 978-93-83083-75-6. [Google Scholar]
  158. Jiao, N.; Robinson, C.; Azam, F.; Thomas, H.; Baltar, F.; Dang, H.; Hardman-Mountford, N.J.; Johnson, M.; Kirchman, D.L.; Koch, B.P.; et al. Mechanisms of Microbial Carbon Sequestration in the Ocean—Future Research Directions. Biogeosciences 2014, 11, 5285–5306. [Google Scholar] [CrossRef]
  159. Ricour, F.; Guidi, L.; Gehlen, M.; DeVries, T.; Legendre, L. Century-Scale Carbon Sequestration Flux throughout the Ocean by the Biological Pump. Nat. Geosci. 2023, 16, 1105–1113. [Google Scholar] [CrossRef]
  160. Jiao, N.; Azam, F. Microbial Carbon Pump and Its Significance for Carbon Sequestration in the Ocean. In Microbial Carbon Pump in the Ocean; Jiao, N., Azam, F., Sanders, S., Eds.; The American Association for the Advancement of Science: Washington, DC, USA, 2011. [Google Scholar]
  161. Dhamu, V.; Qureshi, M.F.; Barckholtz, T.A.; Mhadeshwar, A.B.; Linga, P. Evaluating Liquid CO2 Hydrate Formation Kinetics, Morphology, and Stability in Oceanic Sediments on a Lab Scale Using Top Injection. Chem. Eng. J. 2023, 478, 147200. [Google Scholar] [CrossRef]
  162. Cunningham, A.B.; Gerlach, R.; Spangler, L.; Mitchell, A.C. Microbially Enhanced Geologic Containment of Sequestered Supercritical CO2. Energy Procedia 2009, 1, 3245–3252. [Google Scholar] [CrossRef]
  163. Benson, S.M.; Cole, D.R. CO2 Sequestration in Deep Sedimentary Formations. Elements 2008, 4, 325–331. [Google Scholar] [CrossRef]
  164. Pan, S.-Y.; Chang, E.E.; Chiang, P.-C. CO2 Capture by Accelerated Carbonation of Alkaline Wastes: A Review on Its Principles and Applications. Aerosol Air Qual. Res. 2012, 12, 770–791. [Google Scholar] [CrossRef]
  165. Bodor, M.; Santos, R.; Gerven, T.; Vlad, M. Recent Developments and Perspectives on the Treatment of Industrial Wastes by Mineral Carbonation—A Review. Open Eng. 2013, 3, 566–584. [Google Scholar] [CrossRef]
  166. Gomes, H.I.; Mayes, W.M.; Rogerson, M.; Stewart, D.I.; Burke, I.T. Alkaline Residues and the Environment: A Review of Impacts, Management Practices and Opportunities. J. Clean. Prod. 2016, 112, 3571–3582. [Google Scholar] [CrossRef]
  167. Eloneva, S.; Puheloinen, E.-M.; Kanweva, J.; Ekroos, A.; Zevenhoven, R.; Fogelholm, C.-J. Co-Utilisation of CO2 and Steelmaking Slags for Production of Pure CaCO3—Legislative Issues. J. Clean. Prod. 2010, 18, 1833–1839. [Google Scholar] [CrossRef]
  168. Rashid, M.I.; Yaqoob, Z.; Mujtaba, M.A.; Fayaz, H.; Saleel, C.A. Developments in Mineral Carbonation for Carbon Sequestration. Heliyon 2023, 9, e21796. [Google Scholar] [CrossRef]
  169. Chang, R.; Kim, S.; Lee, S.; Choi, S.; Kim, M.; Park, Y. Calcium Carbonate Precipitation for CO2 Storage and Utilization: A Review of the Carbonate Crystallization and Polymorphism. Front. Energy Res. 2017, 5, 17. [Google Scholar] [CrossRef]
  170. Mayes, W.M.; Younger, P.L. Buffering of Alkaline Steel Slag Leachate across a Natural Wetland. Environ. Sci. Technol. 2006, 40, 1237–1243. [Google Scholar] [CrossRef]
  171. Lim, M.; Han, G.-C.; Ahn, J.-W.; You, K.-S. Environmental Remediation and Conversion of Carbon Dioxide (CO2) into Useful Green Products by Accelerated Carbonation Technology. Int. J. Environ. Res. Public Health 2010, 7, 203–228. [Google Scholar] [CrossRef]
  172. Bobicki, E.R.; Liu, Q.; Xu, Z.; Zeng, H. Carbon Capture and Storage Using Alkaline Industrial Wastes. Prog. Energy Combust. Sci. 2012, 38, 302–320. [Google Scholar] [CrossRef]
  173. Eloneva, S.; Teir, S.; Salminen, J.; Fogelholm, C.-J.; Zevenhoven, R. Fixation of CO2 by Carbonating Calcium Derived from Blast Furnace Slag. Energy 2008, 33, 1461–1467. [Google Scholar] [CrossRef]
  174. Olajire, A.A. A Review of Mineral Carbonation Technology in Sequestration of CO2. J. Pet. Sci. Eng. 2013, 109, 364–392. [Google Scholar] [CrossRef]
  175. Sanna, A.; Uibu, M.; Caramanna, G.; Kuusik, R.; Maroto-Valer, M.M. A Review of Mineral Carbonation Technologies to Sequester CO2. Chem. Soc. Rev. 2014, 43, 8049–8080. [Google Scholar] [CrossRef] [PubMed]
  176. Khudhur, F.W.K.; MacDonald, J.M.; Macente, A.; Daly, L. The Utilization of Alkaline Wastes in Passive Carbon Capture and Sequestration: Promises, Challenges and Environmental Aspects. Sci. Total Environ. 2022, 823, 153553. [Google Scholar] [CrossRef]
  177. Wilson, S.A.; Dipple, G.M.; Power, I.M.; Thom, J.M.; Anderson, R.G.; Raudsepp, M.; Gabites, J.E.; Southam, G. Carbon Dioxide Fixation within Mine Wastes of Ultramafic-Hosted Ore Deposits: Examples from the Clinton Creek and Cassiar Chrysotile Deposits, Canada. Econ. Geol. 2009, 104, 95–112. [Google Scholar] [CrossRef]
  178. Power, I.M.; Harrison, A.L.; Dipple, G.M. Accelerating Mineral Carbonation Using Carbonic Anhydrase. Environ. Sci. Technol. 2016, 50, 2610–2618. [Google Scholar] [CrossRef]
  179. Li, X.; Bertos, M.F.; Hills, C.D.; Carey, P.J.; Simon, S. Accelerated Carbonation of Municipal Solid Waste Incineration Fly Ashes. Waste Manag. 2007, 27, 1200–1206. [Google Scholar] [CrossRef]
  180. Rendek, E.; Ducom, G.; Germain, P. Carbon Dioxide Sequestration in Municipal Solid Waste Incinerator (MSWI) Bottom Ash. J. Hazard. Mater. 2006, 128, 73–79. [Google Scholar] [CrossRef]
  181. Liu, W.; Su, S.; Xu, K.; Chen, Q.; Xu, J.; Sun, Z.; Wang, Y.; Hu, S.; Wang, X.; Xue, Y.; et al. CO2 Sequestration by Direct Gas–Solid Carbonation of Fly Ash with Steam Addition. J. Clean. Prod. 2018, 178, 98–107. [Google Scholar] [CrossRef]
  182. Veetil, S.P.; Pasquier, L.-C.; Blais, J.-F.; Cecchi, E.; Kentish, S.; Mercier, G. Direct Gas–Solid Carbonation of Serpentinite Residues in the Absence and Presence of Water Vapor: A Feasibility Study for Carbon Dioxide Sequestration. Environ. Sci. Pollut. Res. 2015, 22, 13486–13495. [Google Scholar] [CrossRef]
  183. Fernández Bertos, M.; Li, X.; Simons, S.J.R.; Hills, C.D.; Carey, P.J. Investigation of Accelerated Carbonation for the Stabilisation of MSW Incinerator Ashes and the Sequestration of CO2. Green Chem. 2004, 6, 428–436. [Google Scholar] [CrossRef]
  184. Azdarpour, A.; Asadullah, M.; Mohammadian, E.; Hamidi, H.; Junin, R.; Karaei, M.A. A Review on Carbon Dioxide Mineral Carbonation through pH-Swing Process. Chem. Eng. J. 2015, 279, 615–630. [Google Scholar] [CrossRef]
  185. Fagerlund, J.; Nduagu, E.; Romão, I.; Zevenhoven, R. A Stepwise Process for Carbon Dioxide Sequestration Using Magnesium Silicates. Front. Chem. Eng. China 2010, 4, 133–141. [Google Scholar] [CrossRef]
  186. Ben Ghacham, A.; Cecchi, E.; Pasquier, L.-C.; Blais, J.-F.; Mercier, G. CO2 Sequestration Using Waste Concrete and Anorthosite Tailings by Direct Mineral Carbonation in Gas–Solid–Liquid and Gas–Solid Routes. J. Environ. Manag. 2015, 163, 70–77. [Google Scholar] [CrossRef]
  187. El-Naas, M.H.; El Gamal, M.; Hameedi, S.; Mohamed, A.-M.O. CO2 Sequestration Using Accelerated Gas-Solid Carbonation of Pre-Treated EAF Steel-Making Bag House Dust. J. Environ. Manag. 2015, 156, 218–224. [Google Scholar] [CrossRef]
  188. Baciocchi, R.; Polettini, A.; Pomi, R.; Prigiobbe, V.; Von Zedwitz, V.N.; Steinfeld, A. CO2 Sequestration by Direct Gas−Solid Carbonation of Air Pollution Control (APC) Residues. Energy Fuels 2006, 20, 1933–1940. [Google Scholar] [CrossRef]
  189. Ho, H.-J.; Iizuka, A.; Shibata, E.; Tomita, H.; Takano, K.; Endo, T. CO2 Utilization via Direct Aqueous Carbonation of Synthesized Concrete Fines under Atmospheric Pressure. ACS Omega 2020, 5, 15877–15890. [Google Scholar] [CrossRef]
  190. Li, J.; Jacobs, A.D.; Hitch, M. Direct Aqueous Carbonation on Olivine at a CO2 Partial Pressure of 6.5 MPa. Energy 2019, 173, 902–910. [Google Scholar] [CrossRef]
  191. Song, K.; Jang, Y.-N.; Kim, W.; Lee, M.G.; Shin, D.; Bang, J.-H.; Jeon, C.W.; Chae, S.C. Factors Affecting the Precipitation of Pure Calcium Carbonate during the Direct Aqueous Carbonation of Flue Gas Desulfurization Gypsum. Energy 2014, 65, 527–532. [Google Scholar] [CrossRef]
  192. Huijgen, W.J.J.; Witkamp, G.-J.; Comans, R.N.J. Mineral CO2 Sequestration by Steel Slag Carbonation. Environ. Sci. Technol. 2005, 39, 9676–9682. [Google Scholar] [CrossRef]
  193. Yadav, V.S.; Prasad, M.; Khan, J.; Amritphale, S.S.; Singh, M.; Raju, C.B. Sequestration of Carbon Dioxide (CO2) Using Red Mud. J. Hazard. Mater. 2010, 176, 1044–1050. [Google Scholar] [CrossRef] [PubMed]
  194. Uibu, M.; Velts, O.; Kuusik, R. Developments in CO2 Mineral Carbonation of Oil Shale Ash. J. Hazard. Mater. 2010, 174, 209–214. [Google Scholar] [CrossRef] [PubMed]
  195. Montes-Hernandez, G.; Pérez-López, R.; Renard, F.; Nieto, J.M.; Charlet, L. Mineral Sequestration of CO2 by Aqueous Carbonation of Coal Combustion Fly-Ash. J. Hazard. Mater. 2009, 161, 1347–1354. [Google Scholar] [CrossRef]
  196. Li, Z.; Chen, J.; Lv, Z.; Tong, Y.; Ran, J.; Qin, C. Evaluation on Direct Aqueous Carbonation of Industrial/Mining Solid Wastes for CO2 Mineralization. J. Ind. Eng. Chem. 2023, 122, 359–365. [Google Scholar] [CrossRef]
  197. Ho, H.-J.; Iizuka, A.; Shibata, E.; Tomita, H.; Takano, K.; Endo, T. Utilization of CO2 in Direct Aqueous Carbonation of Concrete Fines Generated from Aggregate Recycling: Influences of the Solid–Liquid Ratio and CO2 Concentration. J. Clean. Prod. 2021, 312, 127832. [Google Scholar] [CrossRef]
  198. Ho, H.-J.; Iizuka, A.; Shibata, E. Utilization of Low-Calcium Fly Ash via Direct Aqueous Carbonation with a Low-Energy Input: Determination of Carbonation Reaction and Evaluation of the Potential for CO2 Sequestration and Utilization. J. Environ. Manag. 2021, 288, 112411. [Google Scholar] [CrossRef]
  199. O’Connor, W.K.; Dahlin, D.C.; Nilsen, D.N.; Rush, G.E.; Walters, R.P.; Turner, P.C. CO2 Storage in Solid Form: A Study of Direct Mineral Carbonation; CSIRO: Collinwood, Australia, 2000.
  200. Mun, M.; Cho, H.; Kwon, J. Study on Characteristics of Various Extractants for Mineral Carbonation of Industrial Wastes. J. Environ. Chem. Eng. 2017, 5, 3803–3821. [Google Scholar] [CrossRef]
  201. Park, A.-H.A.; Fan, L.-S. CO2 Mineral Sequestration: Physically Activated Dissolution of Serpentine and pH Swing Process. Chem. Eng. Sci. 2004, 59, 5241–5247. [Google Scholar] [CrossRef]
  202. Wang, X.; Maroto-Valer, M.M. Dissolution of Serpentine Using Recyclable Ammonium Salts for CO2 Mineral Carbonation. Fuel 2011, 90, 1229–1237. [Google Scholar] [CrossRef]
  203. Bosecker, K. Bioleaching: Metal Solubilization by Microorganisms. FEMS Microbiol. Rev. 1997, 20, 591–604. [Google Scholar] [CrossRef]
  204. Schwartzman, D.W.; Volk, T. Biotic Enhancement of Weathering and the Habitability of Earth. Nature 1989, 340, 457–460. [Google Scholar] [CrossRef]
  205. Chiang, Y.W.; Santos, R.M.; Monballiu, A.; Ghyselbrecht, K.; Martens, J.A.; Mattos, M.L.T.; Gerven, T.V.; Meesschaert, B. Effects of Bioleaching on the Chemical, Mineralogical and Morphological Properties of Natural and Waste-Derived Alkaline Materials. Miner. Eng. 2013, 48, 116–125. [Google Scholar] [CrossRef]
  206. Power, I.M.; McCutcheon, J.; Harrison, A.L.; Wilson, S.; Dipple, G.M.; Kelly, S.; Southam, C.; Southam, G. Strategizing Carbon-Neutral Mines: A Case for Pilot Projects. Minerals 2014, 4, 399–436. [Google Scholar] [CrossRef]
  207. Power, I.M.; Dipple, G.M.; Southam, G. Bioleaching of Ultramafic Tailings by Acidithiobacillus Spp. for CO2 Sequestration. Environ. Sci. Technol. 2010, 44, 456–462. [Google Scholar] [CrossRef]
  208. Fang, C.; Achal, V. Enhancing Carbon Neutrality: A Perspective on the Role of Microbially Induced Carbonate Precipitation (MICP). Biogeotechnics 2024, 2, 100083. [Google Scholar] [CrossRef]
  209. Achal, V.; Pan, X.; Fu, Q.; Zhang, D. Biomineralization Based Remediation of As(III) Contaminated Soil by Sporosarcina Ginsengisoli. J. Hazard. Mater. 2012, 201–202, 178–184. [Google Scholar] [CrossRef]
  210. Achal, V.; Pan, X.; Zhang, D. Bioremediation of Strontium (Sr) Contaminated Aquifer Quartz Sand Based on Carbonate Precipitation Induced by Sr Resistant Halomonas sp. Chemosphere 2012, 89, 764–768. [Google Scholar] [CrossRef]
  211. Stabnikov, V.; Jian, C.; Ivanov, V.; Li, Y. Halotolerant, Alkaliphilic Urease-Producing Bacteria from Different Climate Zones and Their Application for Biocementation of Sand. World J. Microbiol. Biotechnol. 2013, 29, 1453–1460. [Google Scholar] [CrossRef]
  212. Duan, Y.; Niu, L.; Li, B.; He, Y.; Xu, X.; Yu, C.; Wang, Z.; Xiao, C.; Zheng, C. Montmorillonite-Coupled Microbially Induced Carbonate Precipitation (MICP) Enhanced Contaminant Removal and Carbon Capture in Cyanide Tailings. J. Environ. Chem. Eng. 2024, 12, 113498. [Google Scholar] [CrossRef]
  213. Achal, V.; Mukherjee, A. A Review of Microbial Precipitation for Sustainable Construction. Constr. Build. Mater. 2015, 93, 1224–1235. [Google Scholar] [CrossRef]
  214. Kang, B.; Zha, F.; Deng, W.; Wang, R.; Sun, X.; Lu, Z. Biocementation of Pyrite Tailings Using Microbially Induced Calcite Carbonate Precipitation. Molecules 2022, 27, 3608. [Google Scholar] [CrossRef] [PubMed]
  215. Maureira, A.; Zapata, M.; Olave, J.; Jeison, D.; Wong, L.-S.; Panico, A.; Hernández, P.; Cisternas, L.A.; Rivas, M. MICP Mediated by Indigenous Bacteria Isolated from Tailings for Biocementation for Reduction of Wind Erosion. Front. Bioeng. Biotechnol. 2024, 12, 1393334. [Google Scholar] [CrossRef] [PubMed]
  216. Mwandira, W.; Nakashima, K.; Kawasaki, S.; Ito, M.; Sato, T.; Igarashi, T.; Banda, K.; Chirwa, M.; Nyambe, I.; Nakayama, S.; et al. Efficacy of Biocementation of Lead Mine Waste from the Kabwe Mine Site Evaluated Using Pararhodobacter sp. Environ. Sci. Pollut. Res. Int. 2019, 26, 15653–15664. [Google Scholar] [CrossRef] [PubMed]
  217. De Muynck, W.; De Belie, N.; Verstraete, W. Microbial Carbonate Precipitation in Construction Materials: A Review. Ecol. Eng. 2010, 36, 118–136. [Google Scholar] [CrossRef]
  218. Bandyopadhyay, A.; Saha, A.; Ghosh, D.; Dam, B.; Samanta, A.K.; Dutta, S. Microbial Repairing of Concrete & Its Role in CO2 Sequestration: A Critical Review. Beni-Suef Univ. J. Basic Appl. Sci. 2023, 12, 7. [Google Scholar] [CrossRef]
  219. Hussain, A.; Ali, D.; Koner, S.; Hseu, Z.-Y.; Hsu, B.-M. Microbial Induce Carbonate Precipitation Derive Bio-Concrete Formation: A Sustainable Solution for Carbon Sequestration and Eco-Friendly Construction. Environ. Res 2025, 270, 121006. [Google Scholar] [CrossRef]
  220. Wong, P.Y.; Mal, J.; Sandak, A.; Luo, L.; Jian, J.; Pradhan, N. Advances in Microbial Self-Healing Concrete: A Critical Review of Mechanisms, Developments, and Future Directions. Sci. Total Environ. 2024, 947, 174553. [Google Scholar] [CrossRef]
  221. Fu, T.; Saracho, A.C.; Haigh, S.K. Microbially Induced Carbonate Precipitation (MICP) for Soil Strengthening: A Comprehensive Review. Biogeotechnics 2023, 1, 100002. [Google Scholar] [CrossRef]
  222. Cheng, L.; Shahin, M.A.; Mujah, D. Influence of Key Environmental Conditions on Microbially Induced Cementation for Soil Stabilization. J. Geotech. Geoenviron. Eng. 2017, 143, 04016083. [Google Scholar] [CrossRef]
  223. DeJong, J.T.; Mortensen, B.M.; Martinez, B.C.; Nelson, D.C. Bio-Mediated Soil Improvement. Ecol. Eng. 2010, 36, 197–210. [Google Scholar] [CrossRef]
  224. Ramanan, R.; Kannan, K.; Sivanesan, S.D.; Mudliar, S.; Kaur, S.; Tripathi, A.K.; Chakrabarti, T. Bio-Sequestration of Carbon Dioxide Using Carbonic Anhydrase Enzyme Purified from Citrobacter Freundii. World J. Microbiol. Biotechnol. 2009, 25, 981–987. [Google Scholar] [CrossRef]
  225. Gilmour, K.A.; Ghimire, P.S.; Wright, J.; Haystead, J.; Dade-Robertson, M.; Zhang, M.; James, P. Microbially Induced Calcium Carbonate Precipitation through CO2 Sequestration via an Engineered Bacillus Subtilis. Microb. Cell Factories 2024, 23, 168. [Google Scholar] [CrossRef] [PubMed]
  226. Li, W.; Chen, W.-S.; Zhou, P.-P.; Yu, L.-J. Influence of Enzyme Concentration on Bio-Sequestration of CO2 in Carbonate Form Using Bacterial Carbonic Anhydrase. Chem. Eng. J. 2013, 232, 149–156. [Google Scholar] [CrossRef]
  227. Silva-Castro, G.A.; Uad, I.; Gonzalez-Martinez, A.; Rivadeneyra, A.; Gonzalez-Lopez, J.; Rivadeneyra, M.A. Bioprecipitation of Calcium Carbonate Crystals by Bacteria Isolated from Saline Environments Grown in Culture Media Amended with Seawater and Real Brine. BioMed Res. Int. 2015, 2015, 816102. [Google Scholar] [CrossRef]
  228. Zheng, T.; Qian, C. Influencing Factors and Formation Mechanism of CaCO3 Precipitation Induced by Microbial Carbonic Anhydrase. Process Biochem. 2020, 91, 271–281. [Google Scholar] [CrossRef]
  229. Xiao, L.; Lian, B. Heterologously Expressed Carbonic Anhydrase from Bacillus mucilaginosus Promoting CaCO3 Formation by Capturing Atmospheric CO2. Carbonates Evaporites 2016, 31, 39–45. [Google Scholar] [CrossRef]
  230. Abdelsamad, R.; Disi, Z.A.; Abu-Dieyeh, M.; Al-Ghouti, M.A.; Zouari, N. Evidencing the Role of Carbonic Anhydrase in the Formation of Carbonate Minerals by Bacterial Strains Isolated from Extreme Environments in Qatar. Heliyon 2022, 8, e11151. [Google Scholar] [CrossRef]
  231. Huang, L.; Li, F.; Ji, C.; Wang, Y.; Yang, G. Carbon Isotope Fractionation and Its Tracer Significance to Carbon Source during Precipitation of Calcium Carbonate in the Presence of Bacillus Cereus LV-1. Chem. Geol. 2022, 609, 121029. [Google Scholar] [CrossRef]
  232. Yang, G.; Li, L.; Li, F.; Zhang, C.; Lyu, J. Mechanism of Carbonate Mineralization Induced by Microbes: Taking Curvibacter Lanceolatus Strain HJ-1 as an Example. Micron 2021, 140, 102980. [Google Scholar] [CrossRef]
  233. Mitchell, A.C.; Dideriksen, K.; Spangler, L.H.; Cunningham, A.B.; Gerlach, R. Microbially Enhanced Carbon Capture and Storage by Mineral-Trapping and Solubility-Trapping. Environ. Sci. Technol. 2010, 44, 5270–5276. [Google Scholar] [CrossRef]
  234. Zhang, Y.; Hu, X.; Wang, Y.; Jiang, N. A Critical Review of Biomineralization in Environmental Geotechnics: Applications, Trends, and Perspectives. Biogeotechnics 2023, 1, 100003. [Google Scholar] [CrossRef]
  235. Government of Canada; Innovation, Science and Economic Development Canada; Cement Association of Canada. Roadmap to Net-Zero Carbon Concrete by 2050; Innovation, Science and Economic Development Canada: Ottawa, ON, Canada, 2024.
  236. del Strother, P. 2—Manufacture of Portland Cement. In Lea’s Chemistry of Cement and Concrete, 15th ed.; Hewlett, P.C., Liska, M., Eds.; Butterworth-Heinemann: Oxford, UK, 2019; pp. 31–56. ISBN 978-0-08-100773-0. [Google Scholar]
  237. Hendriks, C.A.; Worrell, E.; Price, L.; Martin, N.; Ozawa Meida, L.; De Jager, D.; Riemer, P. Emission Reduction of Greenhouse Gases from the Cement Industry. In Greenhouse Gas Control Technologies 4; Elsevier: Amsterdam, The Netherlands, 1999; pp. 939–944. ISBN 978-0-08-043018-8. [Google Scholar]
  238. Sharma, M.; Satyam, N.; Reddy, K.R. State of the Art Review of Emerging and Biogeotechnical Methods for Liquefaction Mitigation in Sands. J. Hazard. Toxic Radioact. Waste 2021, 25, 03120002. [Google Scholar] [CrossRef]
  239. Burbank, M.B.; Weaver, T.J.; Green, T.L.; Williams, B.C.; Crawford, R.L. Precipitation of Calcite by Indigenous Microorganisms to Strengthen Liquefiable Soils. Geomicrobiol. J. 2011, 28, 301–312. [Google Scholar] [CrossRef]
  240. Zhang, K.; Tang, C.-S.; Jiang, N.-J.; Pan, X.-H.; Liu, B.; Wang, Y.-J.; Shi, B. Microbial-induced Carbonate Precipitation (MICP) Technology: A Review on the Fundamentals and Engineering Applications. Environ. Earth Sci. 2023, 82, 229. [Google Scholar] [CrossRef]
  241. Liu, J.; Zhen, B.; Qiu, H.; Zhou, X.; Zhang, H. Impact of Waterlogging and Heat Stress on Rice Rhizosphere Microbiome Assembly and Potential Function in Carbon and Nitrogen Transformation. Arch. Agron. Soil Sci. 2023, 69, 1920–1932. [Google Scholar] [CrossRef]
  242. Liu, B.; Tang, C.-S.; Pan, X.-H.; Zhu, C.; Cheng, Y.-J.; Xu, J.-J.; Shi, B. Potential Drought Mitigation through Microbial Induced Calcite Precipitation—MICP. Water Resour. Res. 2021, 57, e2020WR029434. [Google Scholar] [CrossRef]
Figure 2. Biochemical reactions facilitating the different metabolic pathways of MICP, including redox-driven reactions (red), enzyme-driven reactions (blue), and photosynthesis-driven reactions (green). The reactions involved in the nitrogen (dark red) and sulfur (light red) cycle are highlighted. Adapted from [58,70].
Figure 2. Biochemical reactions facilitating the different metabolic pathways of MICP, including redox-driven reactions (red), enzyme-driven reactions (blue), and photosynthesis-driven reactions (green). The reactions involved in the nitrogen (dark red) and sulfur (light red) cycle are highlighted. Adapted from [58,70].
Ijms 26 02230 g002
Figure 3. Calcium regulation in microorganisms showcasing the interaction and metabolism of calcium, leading to CaCO3 precipitation. Adapted from [52].
Figure 3. Calcium regulation in microorganisms showcasing the interaction and metabolism of calcium, leading to CaCO3 precipitation. Adapted from [52].
Ijms 26 02230 g003
Figure 4. Simplified schematic of the structure-based reaction mechanism of urease. Red indicates the changes step by step. The flap is open (1), and urea enters the activated site, replacing water and binding to carboxyl oxygen (2). The flap closure enables urea binding to Ni2+ (3). The carbon atom on urea undergoes nucleophilic attack via Ni2+ bridging OH, creating a tetrahedral intermediate (4). The Ni bridging OH transfers the hydrogen atom to the distal urea NH2 group, forming NH3+ (5). The distal C-N bond is broken, and all products are released via a flap opening, which rehydrates the active site (6). Adapted from [108,111].
Figure 4. Simplified schematic of the structure-based reaction mechanism of urease. Red indicates the changes step by step. The flap is open (1), and urea enters the activated site, replacing water and binding to carboxyl oxygen (2). The flap closure enables urea binding to Ni2+ (3). The carbon atom on urea undergoes nucleophilic attack via Ni2+ bridging OH, creating a tetrahedral intermediate (4). The Ni bridging OH transfers the hydrogen atom to the distal urea NH2 group, forming NH3+ (5). The distal C-N bond is broken, and all products are released via a flap opening, which rehydrates the active site (6). Adapted from [108,111].
Ijms 26 02230 g004
Figure 6. Simplified schematic of the structure-based reaction mechanism of CA. CA II active site structure (left) and reaction (right). The Zn-bound OH attacks the carboxyl carbon of CO2 (1), creating Zn-bound HCO3 (2). A water molecule (3) replaces the HCO3 bound to Zn (4), and H+ is transferred to solution (5). Adapted from [146,150,151].
Figure 6. Simplified schematic of the structure-based reaction mechanism of CA. CA II active site structure (left) and reaction (right). The Zn-bound OH attacks the carboxyl carbon of CO2 (1), creating Zn-bound HCO3 (2). A water molecule (3) replaces the HCO3 bound to Zn (4), and H+ is transferred to solution (5). Adapted from [146,150,151].
Ijms 26 02230 g006
Figure 7. Direct and indirect carbon sequestration methods. Adapted from [40,41,156].
Figure 7. Direct and indirect carbon sequestration methods. Adapted from [40,41,156].
Ijms 26 02230 g007
Figure 8. Cement production and the GHG emissions associated with specific processes. Adapted from [236,237].
Figure 8. Cement production and the GHG emissions associated with specific processes. Adapted from [236,237].
Ijms 26 02230 g008
Table 3. Direct methods for mineral carbonation and carbon sequestration.
Table 3. Direct methods for mineral carbonation and carbon sequestration.
MethodMaterialCO2
Application
CO2 Input 1Results 1FindingsReference
Direct Gas–Solid CarbonationMunicipal Solid Waste IncineratorCO2 flow100% CO2, 3 bars, 2.5 h3.19% CaCO3 gain bottom ash
7.31% CaCO3 gain fly ash
More suitable to small particle size.[183]
100% CO2, 3 bars, 3 h11% CaCO3 gain fly ashOptimal CO2 capture at water/solid ratio 0.3.[179]
17 bars, 3 h3% CaCO3 gain bottom ashOptimal CO2 capture 20% w/w moisture and 4 mm sieving.[180]
1 bar, 1 h60 g CO2/kg fly ashTemperature (600 °C) and H2O(g) (20%) are more significant than CO2 content.[181]
Waste Concrete & Anorthosite Tailings18.2 vol% CO2, 4 & 5 bar, 30 min66% CO2 removal waste concrete
34% CO2 removal anorthosite
Aqueous phase carbonation resulted in 34.6% removal in 15 min.[186]
Pre-treated EAF steel-making bag house dust3 bar inlet, 1 bar (outlet), 12 L/min0.657 kg CO2/kg dustCarbonation was based on the total calcium content.[187]
Air Pollution Control Residues from a Medical Solid Waste Incinerator100% CO2, 6 h0.12 kg CO2/kg dry solid wasteMaximum carbonation at 400 °C.[188]
Serpentinite Mining ResidueCO2
concentration
18 vol% CO20.07 g CO2/g residueWater vapor (10 vol%) required for carbonation.[182]
Direct Aqueous CarbonationConcrete FinesCO2 flow14% CO2, 90 min0.19 g CO2/g concrete finesAlmost all absorbed CO2 was converted to CaCO3, and increased CO2 concentration requires higher solid–liquid ratio.[189]
Olivine with NaHCO3 & NaOH BufferspCO2 6.5 MPa, 6 h<80% carbonationAgitation is necessary to prevent solids settlement. Low pCO2 requires high NaHCO3 concentration.[190]
Flue Gas Desulfurization Gypsum1 L/min, 15 min90% CaCO3 efficiencyCaCO3 precipitation increased linearly with ammonia content.[191]
Steel Slag19 bar CO2, 30 min0.25 kg CO2/kg steel slagPrimary factors: particle size <2mm to <38 μm and temperature 25–225 °C.[192]
Red-Mud3.5 bar, 3.5 h5.3 g CO2/100 g red mudAt liquid–solid ratio of 0.35.[193]
Oil Shale AshContinuous flow (0.7 m/10 L), 15% CO217–20% bound CO2Size and structure of CaCO3 depended on end-point pH.[194]
Coal Fly Ash10 bars, 18 h26 kg CO2/ton fly-ashPressure was independent of carbonation efficiency and not affected by temperature of fly ash weight.[195]
Industrial/Mining WastesCO2
concentration
15% CO2544.6 g CO2/kg carbide slagCa content in material produces increased carbonation. Max carbon sequestration occurred at < 75 μm particle size, 60 °C, 100 g/L liquid–solid ratio.[196]
Aggregate Recycling Concrete Fines5% CO20.13 g CO2/g concrete fines0.10 CO2/g concrete fines captured as CaCO3, and 0.02 CO2/g concrete fines dissolved in aqueous.[197]
Low-Calcium Fly Ash30% CO20.016 g CO2/g fly ashGood carbonation potential despite low energy input and low calcium content.[198]
1 As reported in the literature.
Table 4. Carbon sequestration using MICP.
Table 4. Carbon sequestration using MICP.
Metabolic PathwayMicrobial StrainMaterialFindingsReference
CACitrobacter freundiiWastewaterCaCO3 precipitated with CO2 catalyzed by CA. Can sequester CO2 at high concentrations, but HCO3 inhibits CA enzyme activity due to pH decrease.[224]
Bacillus subtilisAgar & Liquid MediumCA converted CO2 to CaCO3 minerals.[225]
Bacillus cereusKarst SoilCA enzyme activity influenced CaCO3 crystal morphology.[226]
Bacillus megateriumMortar SpecimensCO2 influx precipitated comparable CaCO3 to ureolysis-precipitated CaCO3[114]
Bacillus pumilus, Bacillus marisflaviSeawater CA   was   observed ,   and   precipitates   included   CaCO 3 · H2O and CaCO3, showing the potential for carbon sequestration.[227]
Bacillus altitudinisMangrove SoilImpact of CO2 sequestration with bacteria showed 75% removal and 97% removal with bacteria and CA.[144]
Bacillus mucilaginosusLiquid MediumOptimal CA at 30 °C and alkaline environment to enhance CO2 hydration.[228]
Bacillus mucilaginosusLiquid MediumCO2 is more easily captured by CA, which alters the size and morphology of CaCO3 crystals.[229]
Psychrobacter sp., Vibrio alginolyticusMarine SedimentsStrong potential for carbonate precipitation with high CA, meaning capture of CO2.[230]
EPS & CABacillus cereusLiquid MediumCalcite induced by bacteria can fix CO2 from air since CO2 released from organic matter is less than in air.[231]
Curvibacter lanceolatusLiquid MediumCA precipitated only calcite, whereas CA and EPA precipitated calcite and aragonite to enhance CO2 fixation.[232]
PhototrophicPhragmoplastophytaDiamond MineSecondary carbonate precipitation capable of offsetting CO2e by 20%.[129]
Oscillatoria sp., Porphyrobacter sp., Blastomas sp., Rhodobacter sp.Diamond MineKimberlite weathering and secondary carbonate precipitation can sequester carbon through photosynthetic bacteria acting as a catalyst to convert CO2 to CaCO3/MgCO3.[130]
UreolysisSporosarcina, Sphingobacterium, Stenotrophomonas, Acinetobacter, ElizabethkingiaCave & Tavern WaterCO2 sequestration depended on pH and the consortia of bacteria.[113]
Sporosarcina, Brevudimonas, Sphingobacterium, Stenotrophomonas, AcinetobacterCave & Tavern WaterAbiotic CO2 sequestration depended on pH and medium, whereas biotic CO2 sequestration depended on the bacterial species or strains.[33]
Sporosarcina pasteuriiTailingsMICP increased CO2 capture from tailings by 27.15–34.55%[212]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wilcox, S.M.; Mulligan, C.N.; Neculita, C.M. Mineral Carbonation for Carbon Sequestration: A Case for MCP and MICP. Int. J. Mol. Sci. 2025, 26, 2230. https://doi.org/10.3390/ijms26052230

AMA Style

Wilcox SM, Mulligan CN, Neculita CM. Mineral Carbonation for Carbon Sequestration: A Case for MCP and MICP. International Journal of Molecular Sciences. 2025; 26(5):2230. https://doi.org/10.3390/ijms26052230

Chicago/Turabian Style

Wilcox, Samantha M., Catherine N. Mulligan, and Carmen Mihaela Neculita. 2025. "Mineral Carbonation for Carbon Sequestration: A Case for MCP and MICP" International Journal of Molecular Sciences 26, no. 5: 2230. https://doi.org/10.3390/ijms26052230

APA Style

Wilcox, S. M., Mulligan, C. N., & Neculita, C. M. (2025). Mineral Carbonation for Carbon Sequestration: A Case for MCP and MICP. International Journal of Molecular Sciences, 26(5), 2230. https://doi.org/10.3390/ijms26052230

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop