Next Article in Journal
Development of Geopolymers from Phosphate By-Products for Thermal Insulation Applications
Previous Article in Journal
Utilization of PMA-EDTC as a Novel Macromolecular Depressant for Galena in the Flotation Separation of Chalcopyrite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Thermodynamic and Kinetic Studies of Dolomite Formation: A Review

by
Chao Chen
1,2,3,
Hanting Zhong
1,2,3,*,
Xinyu Wang
1,4,
Meng Ning
1,2,3,
Xia Wang
1,2,3,
Yuzhu Ge
1,2,3,
Han Wang
1,2,3,
Ruifeng Tang
3,5 and
Mingcai Hou
1,2,3,*
1
State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation, Chengdu University of Technology, Chengdu 610059, China
2
Key Laboratory of Deep-Time Geography & Environment Reconstruction and Applications of Ministry of Natural Resources, Chengdu University of Technology, Chengdu 610059, China
3
Institute of Sedimentary Geology, Chengdu University of Technology, Chengdu 610059, China
4
College of Earth Sciences, Chengdu University of Technology, Chengdu 610059, China
5
Exploration Division of Petrochina Southwest Oil & Gasfield Company, Chengdu 610041, China
*
Authors to whom correspondence should be addressed.
Minerals 2023, 13(12), 1479; https://doi.org/10.3390/min13121479
Submission received: 16 October 2023 / Revised: 19 November 2023 / Accepted: 20 November 2023 / Published: 24 November 2023

Abstract

:
The “dolomite problem”, which has confused scientists for nearly two centuries, is an important fundamental geological problem. The mineralogical characteristics of carbonate minerals show that the dolomite structure consists of an ordered arrangement of alternating layers of Ca2+ and Mg2+ cations interspersed with C O 3 2 anion layers normal to the c-axis. The dolomite structure violates the c glide plane in the calcite structure, which means that dolomite has R 3 ¯ space group symmetry. The ordered dolomite has superlattice XRD reflections [e.g., (101), (015) and (021)], which distinguish it from calcite and high-Mg calcite. The calculation of thermodynamic parameters shows that modern seawater has a thermodynamic tendency of dolomite precipitation and the dolomitization reaction can be carried out in standard state. However, the latest thermodynamic study shows that modern seawater is not conducive to dolomitization, and that seawater is favorable for dolomitization in only a few regions, such as Abu Dhabi, the Mediterranean and the hypersaline lagoons in Brazil. The kinetic factors of dolomite formation mainly consist of the hydration of Mg2+, the presence of sulfate and the activity of carbonate. Current studies have shown that the presence of microorganisms, exopolymeric substances (EPS), organic molecules, carboxyl and hydroxyl functional groups associated with microorganisms and organic molecules, clay minerals with negative charges and dissolved silica facilitate magnesium ions to overcome hydration and thus promote Mg2+ incorporation into growing Ca-Mg carbonates. Similarly, the metabolic activity of microorganisms is conducive to the increase in alkalinity. However, the inhibitory effect of sulfate on dolomite formation seems to be overestimated, and sulfate may even be a catalyst for dolomite formation. Combining the carbonate crystallization mechanism with thermodynamic and kinetic factors suggests that the early stage of dolomite precipitation or the dolomitization reaction may be controlled by kinetics and dominated by unstable intermediate phases, while metastable intermediate phases later transform to ordered dolomite via an Ostwald’s step rule.

1. Introduction

Dolomite [CaMg(CO3)2] is a common carbonate mineral and is the main component of dolostone. Dolomite was first discovered in 1791 by French naturalist Deodal de Dolomieu [1]. Dolostone, an important type of sedimentary rock, can form important oil and gas reservoirs [2], and most Mississippi Valley-type (MVT) Pb-Zn deposits are hosted in dolostone [3], which indicates that dolostone has important economic value.
Dolomite is widely developed in deep-time marine carbonate rocks [4,5]. Although modern seawater is supersaturated with respect to dolomite [6], dolomite is a rare precipitate in Holocene sediments [4,7,8]. Additionally, dolomite is notoriously difficult to precipitate experimentally at ambient temperature and pressure [9]. This is known as the “dolomite problem” [10,11]. Carbonate precipitation is controlled by both kinetic and thermodynamic considerations [12]. Therefore, many scholars have conducted a great deal of research in thermodynamics and kinetics to solve the dolomite problem.
In terms of thermodynamics, the solubility product of dolomite was experimentally determined by scholars as early as the 1950s [13,14,15,16,17,18,19,20], but due to the difficulty in determining equilibrium in laboratory experiments, the values of the dolomite solubility product determined experimentally at ambient temperature have wide discrepancies. Therefore, it is difficult to accurately determine the solubility product of dolomite in standard state. Additionally, many scholars have proposed stoichiometric reactions to form dolomites [21,22,23,24], and some have been used for thermodynamic considerations. At the same time, thermodynamic parameters of related minerals and ions that can be used to determine the direction of the dolomitization reaction have been also published [19,25,26,27,28,29,30,31]. Although these reactions are all in the direction of dolomite precipitation, these reactions have not been demonstrated in the laboratory at low temperatures, and the scarcity of dolomites in modern environments may indicate that reaction kinetics may be a major control in the formation of dolomite [7,9]. In terms of kinetics, it is believed that the main kinetic factors of dolomite precipitation are the hydration of Mg2+ [22], the presence of sulfate [32] and the activity of carbonate [22]. The hydration of Mg2+ is an important factor in inhibiting dolomite precipitation because magnesium ions are tightly bound to water at low temperatures, which makes it difficult for Mg2+ to enter the carbonate lattice. Accordingly, working out how to make magnesium ion dehydrate is the key to solving the dolomite problem. In addition, the presence of sulfate will cause S O 4 2 and Mg2+ in water to form tight ion pairs, thus inhibiting the incorporation of Mg2+ into the dolomite lattice [32]. However, recent studies indicate that the inhibitory effect of sulfate is not strong at low temperatures [33]. Lippmann [22] pointed out that the activity of carbonate in seawater is also a kinetic factor affecting dolomite precipitation. Garrels and Thompson [34] calculated a chemical model for seawater at 25 °C and found that the activity of carbonate (4.7 × 10−6) is much lower than the molality (2.69 × 10−4). Thus, working out how to weaken the kinetic barriers to dolomite formation has become a focus of research in recent years.
Accordingly, in order to better understand the dolomite problem, this paper reviews the research progress in kinetic and thermodynamic studies of dolomite genesis in recent decades. Firstly, this paper reviews the literature on the mineralogy of dolomite in order to distinguish other Ca-Mg carbonate minerals. Secondly, thermodynamic and kinetic studies related to dolomite origin in recent decades are reviewed. Finally, based on the thermodynamic and kinetic considerations of dolomite formation, we discuss the possible formation process of dolomite and propose directions for the study of the genesis of dolomite.

2. Dolomite Mineralogy

2.1. Calcite and Dolomite

When we study the crystal structure of rhombohedral carbonates, the crystal structure of calcite should be considered first. Because some important carbonate mineral constituents of sedimentary rocks, e.g., magnesium- and iron-bearing carbonates, have structures that are identical to or closely related to the crystal structure of calcite [22]. Accordingly, the structure of calcite can be used as a basis for describing the structure of such minerals. Calcite, a rhombohedral carbonate mineral, is a typical trigonal system. The vast majority of diagenetic calcite in the rock record formed from shallow burial in marine-derived fluids [35]. The calcite crystal structure consists of alternating layers of calcium cations (Ca2+) and carbonate anions ( C O 3 2 ) oriented normal to the c-axis (Figure 1a). The calcite crystal structure is of space group R 3 ¯ c [22], and the unit cell of calcite is hexagonal [dimension: a = 4.9896 (Å), c = 17.0610 (Å)] [36,37,38]. The (104) reflection peak is the most intense reflection in calcite and dolomite, which corresponds to the cleavage rhombohedron planes [38], but there are no ‘ordering’ reflections of dolomite in the XRD pattern of calcite.
Dolomite (CaMg(CO3)2) is a special rhombohedral carbonate mineral containing Ca and Mg. Very few sedimentary dolomites are truly stoichiometric and are more suitable to be represented by CaxMg1−x(CO3)2. The stoichiometric compositions of natural dolomite vary from Ca1.16Mg0.84(CO3)2~Ca0.96Mg1.04(CO3)2 [7]. Dolomite is a trigonal system mineral whose crystal structure is a derivative of the calcite crystal structure. Dolomite is distinguished from calcite and other rhombohedral carbonates by its stoichiometry [39], and the ideal dolomite crystal structure consists of alternating layers of Mg2+ and Ca2+ interspersed with C O 3 2 groups oriented normal to the c-axis (Figure 1b) [38]. In addition, the dolomite structure violates the c glide plane in the calcite structure, which means that dolomite has R 3 ¯ symmetry [22]. The hexagonal unit cell dimensions of dolomite are a = 4.8069 (Å), c = 16.0034 (Å) [40,41]. Additionally, single-crystal dolomite is often rhombohedral {10 1 ¯ 1} and sometimes columnar {11 2 ¯ 0} [40]. Analysis of the XRD pattern shows that the crystal structure of dolomite is different from that of calcite. The ordered dolomite has superlattice XRD reflections [e.g., (101), (015) and (021)] (Figure 2a) [5,42,43]. The ratio of the diffraction intensity of the (015) crystal plane to that of the (110) crystal plane is used to calculate the degree of cation ordering in dolomite [δ = r (015)/r (110)] [44]. Ancient sedimentary dolomite tends to have a much more regular lamellar structure than Holocene dolomite. In addition, ancient sedimentary dolomite typically possesses a pervasive modulated microstructure (wavelengths ≈ 200 Å) that is generally parallel to {1014} [7].

2.2. High-Mg Calcite

Calcites containing > 4 mol% MgCO3 are considered to be high-Mg calcite [38]. Liu et al. [45] pointed out that the MgCO3 content of high-Mg calcite ranges from 4 mol% to 36 mol%, while calcite with a MgCO3 content ranging from 36 mol% to 55 mol% is called disordered dolomite. High-Mg calcite is characterized by a large number of Mg2+ cations randomly substituting for Ca2+ cations in a single cation position in the calcite lattice, and the disordered distribution of Ca2+ and Mg2+ in high-Mg calcite is in contrast to the ordered distribution of cations in dolomite. High-Mg calcite usually does not reproduce the ‘ordering’ reflections seen in the XRD pattern of dolomite (Figure 2b), and the (104) diffraction peak position (2θ) of high-Mg calcite is between that of dolomite and calcite [38]. So high-Mg calcite is often rhombohedral [46,47]. At present, it is believed that the genesis of high-Mg calcite is mainly through biological origin [48,49,50,51] and inorganic precipitation in seawater [52]. Biogenic species include red corals [49], red coralline algae [53,54], calcareous sponges [50], sea urchin [55] and so on, and the magnesium contents of these biogenic high-Mg calcites range from 4 mol% to 45 mol% [50]. Moreover, high-Mg calcite is unstable with respect to dolomite and calcite under ambient conditions [56].

2.3. Protodolomite, Very High-Mg Calcite and Disordered Dolomite

Protodolomite, very high-Mg calcite and disordered dolomite are terms used to describe products that have near-dolomite stoichiometry (ca 40 to 50 mol% MgCO3) in experiments. The discussion of these terms has been documented in detail in the review of Gregg et al. [38]. We endorse their view that a rhombohedral Ca-Mg carbonate mineral is dolomite if its XRD pattern does display evidence of ordering reflections. If the mineral does not show evidence of cation ordering, then it is not dolomite and is very high-Mg calcite. This view is also supported by other authors [57,58]. They think the absence of two out of the three principal ordering reflections (the XRD pattern of ideal dolomite has (101), (015) and (021) ‘ordering’ reflections) indicates the arrangement of Ca2+ and Mg2+ in the mineral crystal is not ordered, and it should belong to the calcite space group (R 3 ¯ c) rather than the dolomite space group (R 3 ¯ ).
Over the two decades, the microbial model of dolomite formation has attracted much attention from geologists. Many scholars have carried out simulation experiments in the laboratory on the dolomite precipitation induced by microorganisms at ambient temperature and pressure [59,60,61,62,63,64,65,66]. However, Gregg et al. [38] and Kaczmarek et al. [57] reevaluated the XRD data of the experimental products of microbial-induced experiments and found that the products precipitated by microbial mediation actually lack cation ordering. Therefore, these minerals are not strictly dolomite, but calcite group minerals. They suggested that these products may be very high-Mg calcite that is close to the ideal stoichiometric composition of dolomite (Figure 3). Qiu et al. [66] also found in their experiment using halophilic archaea to mediate dolomite precipitation that XRD and SAED patterns of the products did not have the ordering characteristics of dolomite (Figure 3), and it was likely that disordered dolomite (very high-Mg calcite) formed. The experimental results of Zhang et al. [67] showed that in the presence of ~113 mg L−1 of non-living biomass, disordered dolomite with ~41 and 45 mol% of MgCO3 was precipitated in solutions with initial Mg:Ca ratios of 5:1 and 8:1, respectively. They believed that non-living biomass of the methanogen Methanosarcina barkeri can enhance the incorporation of Mg into the calcitic structure and induce the crystallization of disordered dolomite. Additionally, microbially mediated very high-Mg calcite was observed in coralline algae from Australia [53,68]. These scholars regarded the observed Ca-Mg carbonate minerals (ranging from 38 to 62 mol% MgCO3) as ‘dolomite’ only based on the magnesium content, without considering the cation ordering. Accordingly, the Ca-Mg carbonate minerals these authors observed are probably very high-Mg calcite, not dolomite. Based on the above facts, the products of most microbially induced experiments are still strictly disordered and are more likely to be very high-Mg calcite in nature.

3. Advances in Thermodynamic Studies of Dolomite Origin

Dolomite dissolution can be expressed by the reaction:
CaMg ( CO 3 ) 2 = Ca 2 + + Mg 2 + + 2 C O 3 2 ,
This gives an equilibrium constant Ksp-dol defined by:
K sp-dol = a M g 2 + ( a C O 3 2 ) 2 ,
where Ksp-dol represents the solubility product of dolomite and ai refers to the activity of the subscripted aqueous species at equilibrium [19]. Ideally, the solubility product should be determined using a saturated solution at the desired temperature/pressure conditions [69]. However, even under highly supersaturated conditions, it is difficult to precipitate ordered, stoichiometric dolomite under standard state conditions (25 °C, 1 atm) [7,9]. Therefore, it is difficult to determine the equilibrium in solubility experiments, which makes the values of dolomite solubility products reported in the literature vary widely from 10−16.5 to 10−19.33 at low temperatures. For instance, Yanat’eva [13] reported a Ksp-dol value of 10−18.37 at 25 °C, measured from dolomite dissolution experiments performed in pure water over 100 days. Garrels et al. [14] obtained a Ksp-dol value of 10−19.33 through dolomite dissolution experiments at 25 °C and 1atm pCO2. Hsü [15] assumed that water from the Florida Aquifer was in equilibrium with dolomite and obtained a mean Ksp-dol value of 10−17. Hsü [16] estimated that the stoichiometric dolomite solubility product constant under standard state was ≈10−17, based on modern metastable dolomites. Hardie [17] estimated that the solubility product constant of metastable Ca-rich dolomite was ≈10−16.5. Sherman and Barak [18] performed dissolution experiments with dolomite in Ca-Mg-HCO3/CO3 solutions at 25 °C for 672 days and calculated a Ksp-dol value of 10−17.2±0.2. The solubility product of dolomite was experimentally determined to be 10−17.56~10−26.44 at 50~253 °C using dolomite of hydrothermal origin by Bénézeth et al. [19]. They then fitted a Ksp-dol value at 25 °C of 10−17.19±0.3 from these experimental data. Robertson et al. [20] used groundwater methods instead of experimental methods to determine Ksp-dol. They assumed that calcite-dolomite has reached equilibrium with a large residence time in the subsurface fluid. A Ksp-dol value of 10−17.27±0.35 at 25 °C was then evaluated by a mixed-effects model.
Generally speaking, the direction of the reversible reaction can be determined by comparing the ion product (Q) with the solubility product constant (Ksp) [30]. The ion product is the ion concentration product under any state studied, and the expression is:
Q dol = [ Ca 2 + ]   [ M g 2 + ]   [ C O 3 2 ] 2 ,
where the parentheses [x] indicate ionic concentration. The solubility product constant is the ion product in the reaction equilibrium state. When Ksp-dol = Qdol, it is at equilibrium, and Ksp-dol is a constant under the condition of constant temperature. When Qdol > Ksp-dol, the mineral precipitates. While Qdol < Ksp-dol, the mineral dissolves. The ion activity product of modern seawater is 10−15.01 [7], so modern seawater has a thermodynamic tendency to precipitate dolomite. Unfortunately, dolomite is scarce in modern environments. Accordingly, we have to consider the dolomitization of calcite and/or aragonite.
According to the second law of thermodynamics, the direction and degree of a phase transition between minerals can be described by the total entropy (the sum of the entropy of the system and the entropy of the environment) or the Gibbs free energy. Changes in natural minerals always tend to increase in total entropy or decrease in Gibbs free energy. Gibbs free energy is defined by a change in enthalpy substituted with a change in entropy multiplied by some reference temperature, as shown in the following equation:
∆G = ∆H − T∆S,
where ∆G represents Gibbs free energy change (kJ·mol−1), ∆H is the enthalpy change (kJ·mol−1), ∆S is the entropy change (J·mol−1·K−1) and T (K) is the temperature of each reaction. The change in entropy and change in enthalpy are found to determine the Gibbs free energy of a reaction because Gibbs free energy is a combination of enthalpy change and entropy change.
The change in enthalpy of the reaction in each process is defined by a change in the enthalpy of the product substituted with a change in the enthalpy of reactants as shown in the following equation:
∆H = Hproducts − Hreactants,
The change in the entropy of the reaction is defined by a change in the entropy of the product substituted with a change in the entropy of the reactants as shown in the following equation:
∆S = Sproducts − Sreactants,
Dolomitization of limestone can be expressed by the reaction:
2CaCO3 + Mg2+ = CaMg(CO3)2 + Ca2+,
The reaction is reversible and the direction of the reaction is determined by the Gibbs free energy change (∆rG°). The reaction proceeds positively when ∆rG° < 0. The reaction proceeds in the reverse direction when ∆rG° > 0. The thermodynamic parameters of reactants and products in dolomitization are shown in Table 1. Thermodynamic calculations indicate that calcite can be replaced by dolomite at ordinary temperatures because ∆rG° < 0 (Table 2) for dolomitization in standard state. However, carbonate ions must present in various aquifers and Mg-rich fluids, and not just from the dissolution of calcite or aragonite, as neither calcite nor dolomite can be stable in aqueous solution if there is no dissolved carbonate in the system [70]. Obviously, it is indispensable to consider the effect of C O 3 2 or H C O 3 on the thermodynamic equilibrium. Therefore, reaction (7) may not be suitable for the real dolomitization reaction and thermodynamic equilibrium in the natural environment. Some scholars have proposed other reaction equations for dolomitization. For example, Lippmann [21] proposed a chemical formula for the formation of dolomite based on the synthesis of norsethite (BaMg(CO3)2) at 20 °C:
CaCO 3   + Mg 2 + + C O 3 2 = CaMg ( CO 3 ) 2 ,
The conclusion obtained by the calculation of thermodynamic parameters of reactants and products is consistent with reaction (7), however, reaction (8) requires not only the supply of magnesium but also the supply of carbonate ions. Lippmann [21] explained that the alkalinity required for this process of dolomitization may be derived from the hydrolysis connected with silicate weathering, or it may be produced from magnesium sulfate and organic matter by sulfate-reducing bacteria in buried carbonate sediments. In reaction (7), the dolomitizing fluid supplies the Mg2+ required and removes the Ca2+ released into the fluid; while in reaction (8), the dolomitizing fluid needs to supply Mg2+ and Ca2+ without removing any reaction products. In addition, the enthalpy changes in reactions (7) and (8) are 14.8 kJ·mol−1 and 26.4 kJ·mol−1, respectively, indicating that these two reactions are heat-absorbing and require higher temperatures to proceed spontaneously.
Morrow [23] proposed a dolomitization reaction (reaction (9)) based on the conservation of volume. This reaction is intermediate between reactions (7) and (8).
( 2 x ) CaCO 3 + Mg 2 +   + x C O 3 2 = CaMg ( CO 3 ) 2 + ( 1 x ) Ca 2 +
where x = 0.11 and 0.25 for the dolomitization of aragonite and calcite, respectively. For x = 0.11 and 0.25, reaction (9) is characterized by a free energy change at equilibrium of ∆rG° < 0 (Table 2) for dolomitization of aragonite and calcite. In addition, this reaction can reflect the degree of gain or loss of rock volume during dolomitization [23].
Recently, Wang [24] found that in pure carbonate solution, the saturation exponential model of dolomite cannot fully reflect its precipitation trend due to the inherent relationship between dolomite and calcite and their stoichiometry competition for ions Ca2+ and C O 3 2 . Thereby, he developed a new CDGPE (Calcite-Dolomite-Gypsum Phase Equilibrium) phase equilibrium model as shown in the following formula:
3 CaCO 3 ( s ) + Mg 2 + + S O 4 2 + H + + 2 H 2 O = CaSO 4 · 2 H 2 O ( s ) + CaMg ( CO 3 ) 2 ( s ) + Ca 2 + + H C O 3 ,
The new model uses the phase equilibrium among calcite, dolomite and gypsum to describe the precipitation/dissolution trend of dolomite, and the Gibbs free energy change in reaction (10) is still less than zero. Reaction (10) takes the other ions in solution into account in the thermodynamic equilibrium, which is clearly closer to the natural environment than reactions (7) and (8). This model indicates that the increase in activity ratios of Mg2+ to Ca2+ and S O 4 2 to H C O 3 , as well as the more acidic condition, will promote dolomite precipitate [24]. In addition, this model has proved reliable and useful in predicting the surface water reported in dolomitization and dedolomitization (including in modern seawater) (Figure 4); Modern seawater is located in the dedolomitization zone, and the seawater of Abu Dhabi, the Mediterranean, and the Persian Gulf is located in the dolomitization zone. However, the author did not elaborate on why sulfates or gypsum are so important for dolomite precipitation. Therefore, it is necessary to further clarify the relationship between dolomite and gypsum in future studies because dolomite is often associated with gypsum in ancient strata. Undeniably, this model theoretically provides us with a new perspective on dolomite origin, but it needs to be verified by simulation experiments or natural examples in the future.
Additionally, there is a difference in the Gibbs free energy between ordered and disordered dolomite. The free energy differences due to order–disorder is sizeable, and there may be a more than 1.3 kcal/mol difference in ∆fG° (25 °C, 1 atm) between ordered dolomite and disordered dolomite [7,17,71], which indicates a significant difference of 1.5 orders of magnitude in the solubility constant of dolomite in standard state only due to the degree of cation ordering. Accordingly, completely disordered dolomites (very high-Mg calcite) are the most soluble and least stable, but also the easiest to precipitate at ambient temperatures [7,17].
In summary, from the analysis of thermodynamic data, it can be concluded that: (1) normal seawater has a thermodynamic tendency to precipitate dolomite; (2) calcite and aragonite can be replaced by dolomite at ambient conditions; and (3) the new phase equilibrium model needs to be verified by simulation experiments or natural examples. Nevertheless, the scarcity of modern dolomite leads us to believe that dolomite precipitation may be more controlled by kinetic factors.

4. Advances in Kinetic Studies of Dolomite Origin

Based on the theory of chemical dynamics, in any reaction, not all molecules can take part in the reaction, but molecules (ions) with a certain level of energy can. These molecules are called activated molecules, and the difference between the average energy of activated molecules and the average energy of all molecules is called the activation energy. The reaction first needs to overcome the activation energy to occur. Therefore, even reactions that can proceed spontaneously according to thermodynamic principles may not occur, depending on the energy level of the reacting molecules and the magnitude of the activation energy.
Thermodynamic studies indicate that modern seawater has a thermodynamic tendency to precipitate dolomite. The widespread distribution of dolomite in ancient sedimentary strata also strongly suggests that dolomite is a stable phase in carbonate diagenesis at low temperatures and pressure [16]. However, carbonate rocks precipitated in modern normal marine environments contain little or no dolomite, even if the seawater is supersaturated with dolomite. Similarly, it has so far been difficult to precipitate/synthesize ordered stoichiometric dolomite in the laboratory at low temperatures [9,72]. Many geologists attribute this to kinetic inhibition [7,9,16,73]. To date, the main kinetic factors that inhibit the nucleation or precipitation of dolomite at low temperatures are thought to include: (1) the high hydration energy of the magnesium ion [22,45,51,66,74,75,76,77,78,79,80]; (2) the presence of sulfate (even in very low concentrations), related to the formation of complex MgSO40 [32,74,81,82]; and (3) the low activity of C O 3 2 in most natural solutions [22,70].

4.1. Hydration of the Magnesium Ion

Lippmann [22,83] showed that the supersaturation of dolomite in seawater can last for a long time without precipitating dolomite. He suggested that this reflects the relative strength of the electrostatic bonding of Mg2+ to H2O (greater than that for Ca2+ and C O 3 2 ), i.e., the enthalpy of hydration of Mg2+ is greater than that of Ca2+ (1926 kJ/mole vs. 1579 kJ/mole). Although seawater is supersaturated with dolomite, C O 3 2 cannot overcome the hydrated shell to bond with Mg2+, and Mg2+ is difficult to incorporate into the carbonate mineral lattice. Accordingly, sedimentologists have determined that the hydration of Mg2+ in seawater is an important factor in preventing dolomite precipitation [4,7,22,45,51,64,66,74,75,76,77,78,79,80,84].
In recent years, many researchers have found through experiments factors that can weaken Mg2+ hydration. Very high-Mg calcite with 43 mol% MgCO3 was precipitated by Oomori and Kitano [84] at 40 °C from seawater containing dioxane and Na2CO3. They speculated that the presence of an organic compound reduces the dielectric constant of the solution, thus reducing the hydration of Mg2+ and making it easier to incorporate into the carbonate phase. Zhang et al. [75] showed that the addition of polysaccharides to a solution may help to weaken the chemical bonding between Mg2+ and water molecules, thereby reducing the energy barrier of desolvation of the Mg2+-H2O complex and promoting the bonding of Mg2+ with precipitating carbonate to form very high-Mg calcite. Zhang et al. [76] determined that the addition of a small amount of dissolved sulfide also enhances the bonding of Mg2+ onto the growing carbonate surfaces and reduces the dehydrating energy barrier of the Mg2+-H2O complex, which facilitates the precipitation of very high-Mg calcite. Liu et al. [79] demonstrated through laboratory carbonation experiments that highly negatively charged clay minerals such as illite and montmorillonite can aid the precipitation of abiotic protodolomite under ambient conditions. They speculated that Mg and Ca ions are favorably adsorbed by electrostatic force onto clay surfaces, forming metal–hydroxyl complexes and facilitating the dehydration of Mg2+ and Ca2+ (Figure 5). In addition, Fang and Xu [85] demonstrated experimentally that dissolved silica (Si(OH)4), which has a low dipole moment and dielectric constant, can reduce the dehydration energy barrier of the Mg2+-H2O complex and promote Mg2+ incorporation into the Ca-Mg carbonates.
In the last two decades, some microorganisms, e.g., sulfate-reducing bacteria (SRB) [86,87,88], methanogenic archaea [89,90], aerobic halophilic bacteria [78,91] and anaerobic fermentation bacteria [92], have successfully been used to synthesize Ca-Mg carbonate minerals in the laboratory, which shows that microbial activity can indeed precipitate Ca-Mg carbonate minerals (e.g., disordered dolomite or high magnesium calcite). Some scholars have proposed that the nucleation and growth of Ca-Mg carbonate minerals usually occur on cell walls and/or extracellular polymers (EPS), which not only act as nucleation sites but also facilitate the dehydration of Mg2+ and increase the concentration of available Mg2+, thereby facilitating the formation of high-Mg calcite or disordered dolomite [45,51,65,78,86,88,90,91,92,93,94]. Kenward et al. [64] and Roberts et al. [77] suggested that Mg2+ is complexed and dehydrated by surface-bound carboxyl groups, thereby reducing the energy required for carbonization (i.e., [Mg(H2O)6]2+ + R-COO → [Mg(H2O)5(R-COO)]+ + H2O). Accordingly, natural surfaces with high carboxyl groups (including organic matter and microbial matter) may be a mechanism for the formation of ordered dolomite. Additionally, high-Mg calcite and very high-Mg calcite were precipitated by Liu et al. [45] from a solution containing microbially derived carboxylic acids at 25 °C. They argued that microbially derived carboxylic acids (i.e., citric acid, succinic acid) are effective in diminishing the hydration effect of Mg2+, thereby facilitating the incorporation of Mg2+ into Ca-Mg carbonate minerals.
To summarize, although Mg2+ hydration is an important factor hindering dolomite precipitation, it has been proved in the laboratory that there are many ways to overcome Mg2+ hydration and make Mg2+ incorporate into carbonate minerals to form high-Mg calcite or very high-Mg calcite; and very high-Mg calcite is considered to be a precursor or intermediate product in the dolomite formation [38,57], which shortens the intermediate process of dolomite formation to some extent.

4.2. Sulfate Inhibitor

It is generally believed that S O 4 2 and Mg2+ in water form tight ion pairs, thus inhibiting the incorporation of Mg2+ into the dolomite lattice. This observation originated from Baker and Kastner’s high-temperature (200 °C) synthesis experiment [32]. Later, some scholars also agreed that the presence of sulfate can preclude the formation of dolomite because Mg2+ and S O 4 2 form strong ion pairs [7,61,86,87,95,96,97,98,99,100]. To address this issue, many researchers have suggested that microbial sulfate reduction can mediate/promote dolomite formation by removing sulfate from solution and releasing magnesium ions from strong ion pairs [61,86,87,95,96,97,98,99]. The resolution of the inhibitory effect of dissolved sulfate also facilitated the proposal of the microbial model. However, Sánchez-Román et al. [101] through their culture experiments proved that it is possible to form dolomite when the sulfate concentration was as high as 56 mM. Similarly, the experimental results of Wang et al. [33] demonstrate that at high temperatures, S O 4 2 combines with Mg2+ to form a tight contact ion pair (CIP). Thermochemical sulfate reduction (TSR) can remove sulfate to release Mg2+ and generate CO2 ( C O 3 2 / H C O 3 ), which facilitates the formation of hydrothermal dolomite; whereas at low temperatures, sulfate primarily exists as unassociated S O 4 2 , solvent-separated ion pairs and solvent-shared ion pairs, particularly in very dilute sulfate-bearing solutions (Figure 6). Therefore, they concluded that dissolved sulfate does inhibit the formation of dolomite at high temperatures due to the strong ion pairing between Mg2+ and S O 4 2 but is indeed overestimated at low temperatures. It has also been noted that S O 4 2 may even be a catalyst for dolomite precipitation [102,103], as there are numerous sulphate-rich natural environments and culture experiments that do form very high-Mg calcite and dolomite [8,61,80,86,97,104,105]. The concentration of dissolved sulfate in solution or pore water in areas where dolomite is still precipitating today is still greater than that in seawater (Table 3). Additionally, the symbiosis of dolomite and evaporite in ancient strata is common at different periods of geological history around the world [106,107,108,109,110]. This shows a close correlation between dolomite and gypsum/anhydrite. Thus, dolomitized fluids may be enriched in sulfate. In addition, the thermodynamic phase equilibrium model proposed by Wang [24] suggests that the presence of sulfate is particularly important for dolomite formation; however, this requires further study in simulation experiments and in nature.
In summary, from these current studies, it can be concluded that the removal of sulfate by microbial sulfate reduction at surface temperatures is not the key factor in inhibiting the formation of low-temperature dolomite; rather, sulfate may promote the formation of dolomite [24,102,103].

4.3. The Activity of C O 3 2

Another factor inhibiting dolomite formation is the disproportionate concentrations in seawater of the component ions of dolomite. In particular, C O 3 2 content is extremely low in comparison to the cations [22]. Lippmann [22] argued that only very few carbonate ions have enough kinetic energy to penetrate the hydration barrier due to hydrated Mg2+ on crystal surfaces. The microbial model suggests that the metabolic activity of microorganisms can form C O 3 2 and H C O 3 [e.g., sulfate reduction process Equation (11)], which increases the carbonate alkalinity of the surrounding environment and facilitates the formation of dolomite.
2 CH 2 O + S O 4 2     2 H C O 3 + H 2 S ,

5. Thermodynamic and Kinetic Considerations of Dolomite Nucleation and Growth

For a typical chemical reaction, it is often necessary to distinguish whether it is thermodynamically or kinetically controlled. When the reaction is controlled by kinetics, the reaction system is far from the equilibrium state, and the products are mainly in the transition state or intermediate phase, which can form easily. When the reaction is thermodynamically controlled, the reaction system has reached equilibrium and the products are mainly thermodynamically stable minerals. Thermodynamics deals with the driving force of a system moving from the initial state to the product state, whereas kinetics is concerned with the energy barriers of the specific pathways in this process [113].
Classical nucleation theory suggests that a solid phase forms within a supersaturated solution by adding an atom or a molecule to a growing cluster, whereas the non-classical nucleation theory involves the aggregation of pre-nucleation clusters of ions or primary particles before crystal formation (Figure 7) [114]. Recent studies have shown that a variety of precursor phases exist in the early stages of carbonate crystallization, including pre-nucleation clusters, metastable liquid-like precursors, mesocrystals and amorphous and nanophase precursors [114,115,116,117,118,119,120,121,122,123].
Crystal nucleation and growth are free-energy-driven phenomena and reflect solution supersaturation [121]. The interfacial free energy of the mineral with respect to its solution is a key parameter that determines its nucleation rate. Metastable phases (e.g., (very) high Mg-calcite) with relatively low interfacial free energy and activation energy form at a higher rate than the most stable phases (e.g., ideal dolomite or calcite) [124,125,126].
Previous studies have shown that dolomite precipitation or dolomitization is more likely to be controlled by kinetics. Current studies on the formation of dolomite are also gradually showing the existence of intermediate phases in the formation process [38,72,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143]. For instance, Montes-Hernandez et al. [139] proved that ordered dolomite can precipitate via simultaneous dissolution of calcite and magnesite under hydrothermal conditions (from 100 to 200 °C) (CaCO3 + MgCO3) → CaMg(CO3)2. Two kinetic steps are clearly identified: first, protodolomite formation after about 5 days of reaction; and second, protodolomite to dolomite transformation, probably produced by a coupled dissolution–recrystallization process. Montes-Hernandez et al. [140] showed that Mg-calcite precipitating at ambient temperature (Figure 8a) is quickly dissolved at higher temperatures (200~300 °C) to form dolomite and other Mg-Ca anhydrous carbonates (including protodolomite) (Figure 8b). These metastable mineral phases dissolve in turn to form new dolomite particles or nourish the existing dolomite particles by a crystal growth process. Finally, micron-sized rhombohedrally ordered dolomite particles precipitate (Figure 8c). Rodriguez-Blanco et al. [141] experimentally concluded that the crystallization of dolomite from an aqueous solution at temperatures between 60~220 °C requires three stages. In the first stage, a nanoparticulate magnesium-deficient, amorphous calcium carbonate (Mg-ACC) forms. During the second stage, the Mg-ACC partially dehydrates and orders prior to its rapid (<5 min) crystallization to non-stoichiometric protodolomite (Figure 8d,e). In stage three of the reaction, the protodolomite transforms to highly crystalline and stoichiometric dolomite spheroids on a much longer timescale (hours to days), via an Ostwald-ripening mechanism (~5–50 μm, Figure 8f), but these are made up of larger (>50 and <600 nm) and highly crystalline rhombohedral crystals (Figure 8g). Moreover, metastable very high-Mg calcite (protodolomite) can also transform into dolomite under experimental conditions. For instance, Zheng et al. [5] demonstrated that low-temperature protodolomite (very high-Mg calcite) transforms into dolomite in the absence of external fluid by dry-heating experiments (protodolomites were added into clean quartz tubes filled with ultra-pure nitrogen gas and then sealed under vacuum and heated for two months in the temperature range of 100 °C–300 °C). The superlattice reflections emerge, and the XRD reflections of protodolomite become sharper after protodolomite transforms into dolomite under dry-heating conditions. Furthermore, the products have a more stoichiometric composition (close to 50 mol% MgCO3). The newly formed dolomite retains the spheroidal morphology but exhibits a coarser texture with larger, euhedral nanoscopic grains.
Furthermore, as mentioned in Section 4, the presence of microorganisms, exopolymeric substances (EPS), organic molecules, carboxyl and hydroxyl functional groups associated with microorganisms and organic molecules, clay minerals with negative charges and dissolved silica facilitate the formation of very high-Mg calcite, which is analogous to the reaction paths observed in high-temperature laboratory experiments. Therefore, many scholars believe that very high-Mg calcite is a key precursor mineral during dolomite formation [57,67,79,93,141]. However, this does not indicate that the subsequent conversion of very high-Mg calcite to dolomite must be under high-temperature conditions. It is possible that the process takes a long time under ambient conditions [72,132,144], but this may not be long on a geologic timescale.
The above facts indicate that metastable intermediate phases are easier to generate than thermodynamically stable dolomite during dolomite formation. In the subsequent phase transition stage, it has been suggested that the transformation of high-Mg calcite or very high-Mg calcite to dolomite can be qualitatively described by Ostwald’s step rule [126,138,141,142,143,145,146]. Ostwald’s step rule states that in the course of the transformation of an unstable or metastable system into a stable one (under earth-surface conditions), the system does not go directly to the most stable conformation, but prefers to reach intermediate stages having the closest free energy to the initial state [122], and the process is driven by the surface energy of various crystals [147]. These early-formed, very high-Mg calcites can act as the nuclei for dolomite crystallization during the later dolomitization stage [148]. Metastable intermediate phases recrystallize to ideal dolomite under certain temporal and changing environmental conditions [39]. However, the ordering process of these metastable transition phases has not been studied in detail.
In summary, dolomite nucleation and growth may be consistent with a non-classical crystallization mechanism. Dolomitization proceeds through various disordered phases. The early stage of dolomite precipitation or the dolomitization reaction may be controlled by kinetics and dominated by unstable intermediate phases. These early-formed, very high-Mg calcites transform to dolomite via an Ostwald’s step rule under certain temporal and changing environmental conditions.

6. Conclusions and Prospect

Dolomite [CaMg(CO3)2] is a rhombohedral carbonate mineral. The dolomite structure consists of an ordered arrangement of alternating layers of Ca2+ and Mg2+ cations interspersed with C O 3 2 anion layers normal to the c-axis, which is in contrast to the disordered distribution of Ca2+ and Mg2+ in the (high-Mg) calcite structure.
The thermodynamic characteristics of dolomite formation indicate that modern seawater has a thermodynamic tendency to precipitate dolomite and a thermodynamic drive exists for the conversion of calcite and aragonite to dolomite. However, the latest thermodynamic models predict that modern seawater is not conducive to dolomitization, and that seawater is favorable for dolomitization in only a few regions, such as Abu Dhabi, the Mediterranean and the hypersaline lagoons in Brazil. In terms of kinetics, the presence of microorganisms, exopolymeric substances (EPS), organic molecules, carboxyl and hydroxyl functional groups associated with microorganisms and organic molecules, clay minerals with negative charges and dissolved silica facilitate the incorporation of Mg2+ into growing Ca-Mg carbonates. Microbial metabolic activity also favors an increase in alkalinity and, therefore, dolomite formation. We speculate that the early stage of dolomite precipitation or the dolomitization reaction may be controlled by kinetics and dominated by metastable intermediate phases, while metastable intermediate phases later transform to ordered dolomite via an Ostwald’s step rule.
Finally, since dolomite precipitation or dolomitization reactions are controlled more by kinetics, it is necessary to understand the transition states in simulation experiments and natural environments. The formation process and mechanism of the transformation from metastable high-Mg calcite to dolomite (i.e., ordering process) in low temperatures need to be further explored in the future in the existing small natural salt lakes and evaporative sea areas such as the Lagoa Vermelha lagoon in the east coast of Brazil, Abu Dhabi and in simulation experiments.

Author Contributions

Conceptualization, H.Z. and M.H.; investigation, data curation, C.C. and R.T.; thermomechanical analysis, X.W. (Xinyu Wang); writing—original draft preparation, C.C.; writing—review and editing, C.C., X.W. (Xia Wang), M.N., Y.G., H.W. and H.Z.; project administration, H.Z. and M.H.; funding acquisition, H.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by The National Natural Science Foundation of China (No. 42272131).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

Authors have not received research grants from Petrochina Southwest Oil & Gasfield Company. The authors declare no conflict of interest.

References

  1. Zenger, D.H.; Bourrouilh-Le Jan, F.G.; Carozzi, A.V. Dolomieu and the first description of dolomite. Dolomites A Vol. Honour Dolomieu 1994, 21, 21–28. [Google Scholar]
  2. Zhao, W.Z.; Shen, A.J.; Qiao, Z.F.; Pan, L.Y.; Hu, A.P.; Zhang, J. Genetic types and distinguished characteristics of dolomite and the origin of dolomite reservoirs. Petrol. Explor. Develop. 2018, 45, 983–997. [Google Scholar] [CrossRef]
  3. Li, Q.; Bao, Z.W. Hydrothermal Dolomite: A Review and Perspective. Geotecton. Metallog. 2018, 42, 699–717, (In Chinese with English Abstract). [Google Scholar]
  4. Ning, M.; Huang, K.J.; Shen, B. Applications and advances of the magnesium isotope on the ‘dolomite problem’. Acta Petrol. Sin. 2018, 34, 3690–3708, (In Chinese with English Abstract). [Google Scholar]
  5. Zheng, W.L.; Liu, D.; Yang, S.S.; Fan, Q.G.; Papineau, D.; Wang, H.M.; Qiu, X.; Chang, B.; She, Z.B. Transformation of protodolomite to dolomite proceeds under dry-heating conditions. Earth Planet. Sci. Lett. 2021, 576, 117249. [Google Scholar] [CrossRef]
  6. McKenzie, J.A.; Vasconcelos, C. Dolomite Mountains and the origin of the dolomite rock of which they mainly consist: Historical developments and new perspectives. Sedimentology 2009, 56, 205–219. [Google Scholar] [CrossRef]
  7. Warren, J. Dolomite: Occurrence, evolution and economically important associations. Earth Sci. Rev. 2000, 52, 1–81. [Google Scholar] [CrossRef]
  8. Cheng, J.R.; Meng, X.Q.; Zhang, E.L.; Jiang, Q.F.; Ni, Z.Y.; Ji, J.F. An Early Holocene Primary Dolomite Layer of Abiotic Origin in Lake Sayram, Central Asia. Geophys. Res. Lett. 2021, 48, 096309. [Google Scholar] [CrossRef]
  9. Land, L.S. Failure to precipitate dolomite at 25 °C from dilute solution despite 1000-fold oversaturation after 32 years. Aquat. Geochem. 1998, 4, 361–368. [Google Scholar] [CrossRef]
  10. Fairbridge, R.W. The Dolomite Question; Special Publication: New York, NY, USA, 1957; pp. 125–178. [Google Scholar]
  11. Machel, H.G. Concepts and models of dolomitization: A critical reappraisal. Geol. Soc. Lond. Spec. Publ. 2004, 235, 7–63. [Google Scholar] [CrossRef]
  12. Folk, R.L.; Land, L.S. Mg/Ca ratio and salinity: Two controls over crystallization of dolomite. Am. Assoc. Petrol. Geol. Bull. 1975, 59, 60–68. [Google Scholar]
  13. Yanat’eva, O.K. Solubility of dolomite in water salt solutions. Izv. Sekt. FkhA Akad. Nauk. SSSR 1952, 20, 252–268. (In Russian) [Google Scholar]
  14. Garrels, R.; Thompson, M.E.; Siever, R. Stability of some carbonates at 25 °C and one atmosphere total pressure. Am. J. Sci. 1960, 258, 402–418. [Google Scholar] [CrossRef]
  15. Hsü, K.J. Solubility of dolomite and composition of Florida groundwaters. J. Hydrol. 1963, 1, 288–310. [Google Scholar] [CrossRef]
  16. Hsü, K.J. Chemistry of dolomite formation. Dev. Sediment. 1967, 9, 169–191. [Google Scholar]
  17. Hardie, L.A. Dolomitization; a critical view of some current views. J. Sed. Petrol. 1987, 57, 166–183. [Google Scholar] [CrossRef]
  18. Sherman, L.A.; Barak, P. Solubility and dissolution kinetics of dolomite in Ca-Mg-HCO3/CO3 solutions at 25 °C and 0.1 MPa carbon dioxide. Soil Sci. Soc. Am. J. 2000, 64, 1959–1968. [Google Scholar] [CrossRef]
  19. Bénézeth, P.; Berninger, U.N.; Bovet, N.; Schott, J.; Oelkers, E.H. Experimental determination of the solubility product of dolomite at 50–253 °C. Geochim. Cosmochim. Acta 2018, 224, 262–275. [Google Scholar] [CrossRef]
  20. Robertson, H.A.; Corlett, H.; Hollis, C.; Whitaker, F.F. Solubility product constants for natural dolomite (0–200 °C) through a groundwater-based approach using the USGS produced water database. Am. J. Sci. 2022, 322, 593–645. [Google Scholar] [CrossRef] [PubMed]
  21. Lippmann, F. Synthesis of BaMg(CO3)2 (Norsethite) at 20 °C and the formation of dolomite in sediments. In Recent Developments in Carbonate Sedimentology in Central Europe; Springer Verlag: New York, NY, USA, 1968; pp. 33–37. [Google Scholar]
  22. Lippmann, F. Sedimentary Carbonate Minerals; Springer: Berlin/Heidelberg, Germany, 1973; 228p. [Google Scholar]
  23. Morrow, D.W. Diagenesis I. Dolomite-part I: The chemistry of dolomitization and dolomite precipitation. Geosci. Can. 1982, 9, 5–13. [Google Scholar]
  24. Wang, X.Y. Equilibrium between dolomitization and dedolomitization of a global set of surface water samples: A new theoretical insight on the dolomite inorganic formation mechanism. Mar. Chem. 2021, 235, 104017. [Google Scholar] [CrossRef]
  25. Berman, R.G. Internally consistent thermodynamic data for minerals in the system Na2O-K2O-CaO-MgO-FeO-Fe2O3-Al2O3-SiO2-TiO2-H2O-CO2. J. Petrol. 1988, 29, 445–522. [Google Scholar] [CrossRef]
  26. Hemingway, B.S.; Robie, R.A. Enthalpy and Gibbs Energy of Formation of Dolomite, CaMg(CO3)2, at 298.15 K from HCl Solution Calorimetry; Open-File Report; U.S. Geological Survey: Reston, VA, USA, 1994; pp. 94–575.
  27. Shock, E.L.; Sassani, D.C.; Willis, M.; Sverjensky, D.A. Inorganic species in geologic fluids: Correlations among standard molal thermodynamic properties of aqueous ions and hydroxide complexes. Geochim. Cosmochim. Acta 1997, 61, 907–950. [Google Scholar] [CrossRef] [PubMed]
  28. Holland, T.J.B.; Powell, R. An internally consistent thermodynamic data set for phases of petrological interest. J. Metamorph. Geol. 1998, 16, 309–343. [Google Scholar] [CrossRef]
  29. Rock, P.A.; Mandell, G.K.; Casey, W.H.; Walling, E.M. Gibbs energy of formation of dolomite from electrochemical cell measurements and theoretical calculations. Am. J. Sci. 2001, 301, 103–111. [Google Scholar] [CrossRef]
  30. Chen, Y.Q.; Zhou, X.Y.; Yang, W.J. Thermodynamic and Kinetic Principle in Dolomitizition Process and the Division of Dolomistone Diagenetic Environment. Mar. Orig. Pet. Geol. 2009, 14, 21–25, (In Chinese with English Abstract). [Google Scholar]
  31. Nordstrom, D.K. Improving internal consistency of standard state thermodynamic data for sulfate ion, portlandite, gypsum, barite, celestine, and associated ions. Procedia Earth Planet. Sci. 2013, 7, 624–627. [Google Scholar] [CrossRef]
  32. Baker, E.R.; Kastner, M. Constraints on the formation of sedimentary dolomite. Science 1981, 213, 214–216. [Google Scholar] [CrossRef]
  33. Wang, X.L.; Chou, I.M.; Hu, W.X.; Yuan, S.D.; Liu, H.; Wan, Y.; Wang, X.Y. Kinetic inhibition of dolomite precipitation: Insights from Raman spectroscopy of Mg2+-SO42− ion pairing in MgSO4/MgCl2/NaCl solutions at temperatures of 25 to 200 °C. Chem. Geol. 2016, 435, 10–21. [Google Scholar] [CrossRef]
  34. Garrels, R.; Thompson, M.E. A chemical model for sea water at 25 °C and one atomosphere total pressure. Am. J. Sci. 1962, 260, 57–66. [Google Scholar] [CrossRef]
  35. Hasiuk, F.J.; Kaczmarek, S.E.; Fullmer, S.M. Diagenetic origins of the calcite microcrystals that host microporosity in limestone reservoirs. J. Sediment. Res. 2016, 86, 1163–1178. [Google Scholar] [CrossRef]
  36. Graf, D.L. Crystallographic tables for the rhombohedral carbonates. Am. Mineral. 1961, 46, 1283–1316. [Google Scholar]
  37. Liu, J.Y.; Wang, Z.Y. Crystal structure characterization and X-ray study of dolomite. Mineral. Petrol. 1988, 8, 28–33, (In Chinese with English Abstract). [Google Scholar]
  38. 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]
  39. Tribble, J.S.; Arvidson, R.S.; Lane, M., III; Mackenzie, F.T. Crystal chemistry, and thermodynamic and kinetic properties of calcite, dolomite, apatite, and biogenic silica: Applications to petrologic problems. Sediment. Geol. 1995, 95, 11–37. [Google Scholar] [CrossRef]
  40. Yang, S.Y.; Huang, Z.C.; Chen, Z.N. Microstructural Correlation between Primary and Replacement Dolomites. Acta Petrol. Mineral. 1997, 16, 362–366, (In Chinese with English Abstract). [Google Scholar]
  41. Zhang, J.; Shou, J.F.; Zhang, T.F.; Pan, L.Y.; Zhou, J.G. New Approach on the Study of Dolomite Origin: The crystal structure analysis of dolomite. Acta Sedimentol. Sin. 2014, 32, 550–559, (In Chinese with English Abstract). [Google Scholar]
  42. Riechelmann, S.; Mavromatis, V.; Buhl, D.; Dietzel, M.; Immenhauser, A. Controls on formation and alteration of early diagenetic dolomite: A multi-proxy δ44/40Ca, δ26Mg, δ18O and δ13C approach. Geochim. Cosmochim. Acta 2020, 283, 167–183. [Google Scholar] [CrossRef]
  43. Cai, W.K.; Liu, J.H.; Zhou, C.H.; Keeling, J.; Glasmacher, U.A. Structure, genesis and resources efficiency of dolomite: New insights and remaining enigmas. Chem. Geol. 2021, 573, 120191. [Google Scholar] [CrossRef]
  44. Huang, Z.C.; Yang, S.Y.; Chen, Z.N. Mineralogical study on primary dolomite and Replacement dolomite. Sci. China Ser. D 1996, 26, 544–550, (In Chinese with English Abstract). [Google Scholar]
  45. Liu, D.; Xu, Y.Y.; Yu, Q.Q.; Yu, N.; Qiu, X.; Wang, H.M.; Papineau, D. Catalytic effect of microbially-derived carboxylic acids on the precipitation of Mg-calcite and disordered dolomite: Implications for sedimentary dolomite formation. J. Asian Earth Sci. 2020, 193, 104301. [Google Scholar] [CrossRef]
  46. Ge, Y.Z.; Pederson, C.L.; Lokier, S.W.; Traas, J.P.; Nehrke, G.; Neuser, R.D.; Goetschl, K.E.; Immenhauser, A. Late Holocene to Recent aragonite-cemented transgressive lag deposits in the Abu Dhabi lagoon and intertidal sabkha. Sedimentology 2020, 67, 2426–2454. [Google Scholar] [CrossRef]
  47. Goetschl, K.E.; Dietzel, M.; Purgstaller, B.; Crengg, C.; Mavromatis, V. Control of MgSO40(aq) on the transformation of amorphous calcium carbonate to high-Mg calcite and long-term reactivity of the crystalline solid. Geochim. Cosmochim. Acta 2021, 312, 357–374. [Google Scholar] [CrossRef]
  48. Boussetta, S.; Bassinot, F.; Sabbatini, A.; Caillon, N.; Nouet, J.; Kallel, N.; Rebaubier, H.; Klinkhammer, G.; Labeyrie, L. Diagenetic Mg-rich calcite in Mediterranean sediments: Quantification and impact on foraminiferal Mg/Ca thermometry. Mar. Geol. 2011, 280, 195–204. [Google Scholar] [CrossRef]
  49. Floquet, N.; Vielzeuf, D. Ordered misorientations and preferential directions of growth in mesocrystalline red coral sclerites. Cryst. Growth Des. 2012, 12, 4805–4820. [Google Scholar] [CrossRef]
  50. Long, X.; Ma, Y.R.; Qi, L.M. Biogenic and synthetic high magnesium calcite—A review. J. Struct Biol. 2014, 185, 1–14. [Google Scholar] [CrossRef] [PubMed]
  51. Al Disi, Z.A.; Zouari, N.; Dittrich, M.; Jaoua, S.; Al-Kuwari, H.A.S.; Bontognali, T.R.R. Characterization of the extracellular polymeric substances (EPS) of Virgibacillus strains capable of mediating the formation of high Mg-calcite and protodolomite. Mar. Chem. 2019, 216, 103693. [Google Scholar] [CrossRef]
  52. Malta, J.V.; Castro, J.W.A.; Cabral, C.L.; Fernandes, D.; Cawthra, H.C. Genesis and age of beachrocks on the Rio de Janeiro coastline, Southeast–Brazil. Mar. Geol. 2021, 442, 106649. [Google Scholar] [CrossRef]
  53. Nash, M.C.; Troitzsch, U.; Opdyke, B.N.; Trafford, J.M.; Russell, B.D.; Kline, D.I. First discovery of dolomite and magnesite in living coralline algae and its geobiological implications. Biogeosciences 2011, 8, 3331–3340. [Google Scholar] [CrossRef]
  54. Nash, M.C.; Adey, W.; Harvey, A.S. High magnesium calcite and dolomite composition carbonate in Amphiroa (Lithophyllaceae, Corallinales, Rhodophyta): Further documentation of elevated Mg in Corallinales with climate change implications. J. Phycol. 2021, 57, 496–509. [Google Scholar] [CrossRef]
  55. Veis, A. Organic matrix-related mineralization of sea urchin spicules, spines, test and teeth. Front. Biosci. 2011, 16, 2540–2560. [Google Scholar] [CrossRef] [PubMed]
  56. Land, L.S. Dolomitization; AAPG: Tulsa, OK, USA, 1982; pp. 1–20. [Google Scholar]
  57. Kaczmarek, S.E.; Gregg, J.M.; Bish, D.L.; Machel, H.G.; Fouke, B.W. Dolomite, very-high magnesium calcite, and microbes—Implications for the microbial model of dolomitization. In Characterization and Modeling of Carbonates–Mountjoy Symposium 1; SPEM Special Publication: New York, NY, USA, 2017; Volume 109, pp. 7–20. [Google Scholar]
  58. Zhao, D.F.; Tan, X.C.; Luo, B.; Wang, X.F.; Qiao, Z.F.; Luo, S.Q. A review of Microbial Dolomite: Advances and challenges. Acta Sedimentol. Sin. 2022, 40, 335–349, (In Chinese with English Abstract). [Google Scholar]
  59. Vasconcelos, C.; Mckenzie, J.A.; Bernasconi, S.; Grujic, D.; Tiens, A.J. Microbial mediation as a possible mechanism for natural dolomite formation at low temperatures. Nature 1995, 377, 220–222. [Google Scholar] [CrossRef]
  60. Vasconcelos, C.; McKenzie, J.A.; Warthmann, R.; Bernasconi, S.M. Calibration of the δ18O paleothermometer for dolomite precipitated in microbial cultures and natural environments. Geology 2005, 33, 317–320. [Google Scholar] [CrossRef]
  61. Wright, D.T.; Wacey, D. Precipitation of dolomite using sulphate-reducing bacteria from the Coorong region, South Australia: Significance and implications. Sedimentology 2005, 52, 987–1008. [Google Scholar] [CrossRef]
  62. Sánchez-Román, M.; Vasconcelos, C.; Schmid, T.; Dittrich, M.; McKenzie, J.A.; Zenobi, R.; Rivadeneyra, M.A. Aerobic microbial dolomite at the nanometer scale: Implications for the geologic record. Geology 2008, 36, 879–882. [Google Scholar] [CrossRef]
  63. Sánchez-Román, M.; McKenzie, J.A.; Wagener, A.D.L.R.; Romanek, C.S.; Sánchez-Navas, A.; Vasconcelos, C. Experimentally determined biomediated Sr partition coefficient for dolomite: Significance and implication for natural dolomite. Geochim. Cosmochim. Acta 2011, 75, 887–904. [Google Scholar] [CrossRef]
  64. Kenward, P.A.; Fowle, D.A.; Goldstein, R.H.; Ueshima, M.; González, L.A.; Roberts, J.A. Ordered low temperature dolomite mediated by carboxyl-group density of microbial cell walls. AAPG Bull. 2013, 97, 2113–2125. [Google Scholar] [CrossRef]
  65. Bontognali, T.R.R.; Mckenzie, J.A.; Warthman, R.J.; Vasconcelos, C. Microbially influenced formation of Mg-calcite and Ca-dolomite in the presence of exopolymeric substances produced by sulphate-reducing bacteria. Terra Nova 2014, 26, 72–77. [Google Scholar] [CrossRef]
  66. Qiu, X.; Wang, H.M.; Yao, Y.C.; Duan, Y. High salinity facilitates dolomite precipitation mediated by Haloferax volcanii DS52. Earth Planet. Sci. Lett. 2017, 472, 197–205. [Google Scholar] [CrossRef]
  67. Zhang, F.F.; Xu, H.F.; Shelobolina, E.S.; Konishi, H.; Roden, E.E. Precipitation of low-temperature disordered dolomite induced by extracellular polymeric substances of methanogenic Archaea Methanosarcina barkeri: Implications for sedimentary dolomite formation. Am. Mineral. 2021, 106, 69–81. [Google Scholar] [CrossRef]
  68. Diaz-Pulido, G.; Nash, M.C.; Anthony, K.R.N.; Bender, D.; Opdyke, B.N.; Reyes-Nivia, C.; Troitzsch, U. Greenhouse conditions induce mineralogical changes and dolomite accumulation in coralline algae on tropical reefs. Nat. Commun. 2014, 5, 3310. [Google Scholar] [CrossRef] [PubMed]
  69. Hefter, G.T.; Tomkins, R.P. (Eds.) The Experimental Determination of Solubilities; John Wiley & Sons: Hoboken, NJ, USA, 2003. [Google Scholar]
  70. Machel, H.G.; Mountjoy, E.W. Chemistry and Environments of Dolomitization—A Reappraisal. Earth Sci. Rev. 1986, 23, 175–222. [Google Scholar] [CrossRef]
  71. Helgeson, H.C. Summary and Critique of the Thermodynamic Properties of Rock-Forming Minerals. Am. J. Sci. 1978, 278, 1–229. [Google Scholar]
  72. Usdowski, E. Synthesis of dolomite and magnesite at 60 °C in the system Ca2+-Mg2+-CO32–-Cl22–-H2O. Naturwissenchaften 1989, 76, 374–375. [Google Scholar] [CrossRef]
  73. Land, L.S. The origin of massive dolomite. J. Geol. Educ. 1985, 33, 112–125. [Google Scholar] [CrossRef]
  74. Slaughter, M.; Hill, R.J. The influence of organic matter in organogenic dolomitization. J. Sediment. Res. 1991, 61, 296–303. [Google Scholar] [CrossRef]
  75. Zhang, F.F.; Xu, H.F.; Konishi, H.; Shelobolina, E.S.; Roden, E.E. Polysaccharide-catalyzed nucleation and growth of disordered dolomite: A potential precursor of sedimentary dolomite. Am. Mineral. 2012, 97, 556–567. [Google Scholar] [CrossRef]
  76. Zhang, F.F.; Xu, H.F.; Konishi, H.; Kenp, J.M.; Roden, E.E.; Shen, Z.Z. Dissolved sulfide-catalyzed precipitation of disordered dolomite: Implications for the formation mechanism of sedimentary dolomite. Geochim. Cosmochim. Acta 2012, 97, 148–165. [Google Scholar] [CrossRef]
  77. Roberts, J.A.; Kenward, P.A.; Fowle, D.A.; Goldstein, R.H.; González, L.A.; Moore, D.S. Surface chemistry allows for abiotic precipitation of dolomite at low temperature. Proc. Natl. Acad. Sci. USA 2013, 110, 14540–14545. [Google Scholar] [CrossRef]
  78. Huang, Y.R.; Yao, Q.Z.; Wang, F.P.; Zhou, G.T.; Fu, S.Q. Aerobically incubated bacterial biomass-promoted formation of disordered dolomite and implication for dolomite formation. Chem. Geol. 2019, 523, 19–30. [Google Scholar] [CrossRef]
  79. Liu, D.; Xu, Y.Y.; Papineau, D.; Yu, N.; Fan, Q.; Qiu, X.; Wang, H.M. Experimental evidence for abiotic formation of low-temperature proto-dolomite facilitated by clay minerals. Geochim. Cosmochim. Acta 2019, 247, 83–95. [Google Scholar] [CrossRef]
  80. Liu, D.; Yu, N.; Papineau, D.; Fan, Q.; Wang, H.M.; Qiu, X.; She, Z.B.; Luo, G.M. The catalytic role of planktonic aerobic heterotrophic bacteria in protodolomite formation: Results from Lake Jibuhulangtu Nuur, Inner Mongolia, China. Geochim. Cosmochim. Acta 2019, 263, 31–49. [Google Scholar] [CrossRef]
  81. Kastner, M. Control of dolomite formation. Nature 1984, 311, 410–411. [Google Scholar] [CrossRef]
  82. Morrow, D.W.; Ricketts, B.D. Experimental investigation of sulfate inhibition of dolomite and its mineral analogues. SEPM 1988, 43, 25–38. [Google Scholar]
  83. Lippmann, F. Stable and metastable solubility diagrams for the system CaCO3-MgCO3-H2O at ordinary temperature. Bull. Mineral. 1982, 105, 273–279. [Google Scholar] [CrossRef]
  84. Oomori, T.; Kitano, Y. Synthesis of protodolomite from sea water containing dioxin. Geochem. J. 1987, 21, 59–65. [Google Scholar] [CrossRef]
  85. Fang, Y.H.; Xu, H.F. Dissolved silica-catalyzed disordered dolomite precipitation. Am. Mineral. 2022, 107, 443–452. [Google Scholar] [CrossRef]
  86. Warthmann, R.; van Lith, Y.; Vasconcelos, C.; McKenzie, J.A.; Karpoff, A.M. Bacterially induced dolomite precipitation in anoxic culture experiments. Geology 2000, 28, 1091–1094. [Google Scholar] [CrossRef]
  87. Van Lith, Y.; Warthmann, R.; Vasconcelos, C.; Mckenzie, J.A. Sulphate-reducing bacteria induce low-temperature Ca-dolomite and high Mg-calcite formation. Geobiology 2003, 1, 71–79. [Google Scholar] [CrossRef]
  88. Krause, S.; Liebetrau, V.; Gorb, S.; Sánchez-Román, M.; McKenzie, J.A.; Treude, T. Microbial nucleation of Mg-rich dolomite in exopolymeric substances under anoxic modern seawater salinity: New insight into an old enigma. Geology 2012, 40, 587–590. [Google Scholar] [CrossRef]
  89. Roberts, J.A.; Bennett, P.C.; González, L.A.; Macpherson, G.L.; Milliken, K.L. Microbial precipitation of dolomite in methanogenic groundwater. Geology 2004, 32, 277–280. [Google Scholar] [CrossRef]
  90. Kenward, P.A.; Goldstein, R.H.; Gonzalez, L.A.; Roberts, J.A. Precipitation of low-temperature dolomite from an anaerobic microbial consortium: The role of methanogenic Archaea. Geobiology 2009, 7, 556–565. [Google Scholar] [CrossRef] [PubMed]
  91. Sánchez-Román, M.; Romanek, C.S.; Fernández-Remolar, D.C.; Sánchez-Navas, A.; McKenzie, J.A.; Pibernat, R.A.; Vasconcelos, C. Aerobic biomineralization of Mg-rich carbonates: Implications for natural environments. Chem. Geol. 2011, 281, 143–150. [Google Scholar] [CrossRef]
  92. Zhang, F.F.; Xu, H.F.; Shelobolina, E.S.; Konishi, H.; Converse, B.; Shen, Z.Z.; Roden, E.E. The catalytic effect of bound extracellular polymeric substances excreted by anaerobic microorganisms on Ca-Mg carbonate precipitation: Implications for the “dolomite problem”. Am. Mineral. 2015, 100, 483–494. [Google Scholar] [CrossRef]
  93. Liu, D.; Fan, Q.G.; Papineau, D.; Yu, N.; Chu, Y.Y.; Wang, H.M.; Qiu, X.; Wang, X.J. Precipitation of protodolomite facilitated by sulfate-reducing bacteria: The role of capsule extracellular polymeric substances. Chem. Geol. 2020, 533, 119415. [Google Scholar] [CrossRef]
  94. Diloreto, Z.A.; Garg, S.; Bontognali, T.R.; Dittrich, M. Modern dolomite formation caused by seasonal cycling of oxygenic phototrophs and anoxygenic phototrophs in a hypersaline sabkha. Sci. Rep. 2021, 11, 4170. [Google Scholar] [CrossRef]
  95. Vasconcelos, C.; McKenzie, J.A. Microbial mediation of modern dolomite precipitation and diagenesis under anoxic conditions (Lagoa Vermalha, Rio De Janerio, Brazil). J. Sed. Res. 1997, 67, 378–390. [Google Scholar]
  96. Wright, D.T. The role of sulphate-reducing bacteria and cyanobacteria in dolomite formation in distal ephemeral lakes of the Coorong region, South Australia. Sediment. Geol. 1999, 126, 147–157. [Google Scholar] [CrossRef]
  97. Van Lith, Y.; Vasconcelos, C.; Warthmann, R.; Martins, J.C.F.; McKenzie, J.A. Bacterial sulfate reduction and salinity: Two controls on dolomite precipitation in Lagoa Vermelha and Brejo do Espinho (Brazil). Hydrobiologia 2002, 485, 25–49. [Google Scholar] [CrossRef]
  98. Van Lith, Y.; Warthmann, R.; Vasconcelos, C.; McKenzie, J.A. Microbial fossilization in carbonate sediments: A result of the bacterial surface involvement in dolomite precipitation. Sedimentology 2003, 50, 237–245. [Google Scholar] [CrossRef]
  99. Wright, D.T.; Wacey, D. Sedimentary dolomite: A reality check. In: Braithwaite, C.J.R., Rizzi, G., Darke, G. The Geometry and Petrogenesis of Dolomite Hydrocarbon Reservoirs. Geol. Soc. Lond. Spec. Publ. 2004, 235, 65–74. [Google Scholar] [CrossRef]
  100. Baldermann, A.; Deditius, A.P.; Dietzel, M.; Fichtner, V.; Fischer, C.; Hippler, D.; Leis, A.; Baldermann, C.; Mavromatis, V.; Stickler, C.P.; et al. The role of bacterial sulfate reduction during dolomite precipitation: Implications from Upper Jurassic platform carbonates. Chem. Geol. 2015, 412, 1–14. [Google Scholar] [CrossRef]
  101. Sánchez-Román, M.; McKenzie, J.A.; Wagener, A.D.L.R.; Rivadeneyra, M.A.; Vasconcelos, C. Presence of sulfate does not inhibit low-temperature dolomite precipitation. Earth Planet. Sci. Lett. 2009, 285, 131–139. [Google Scholar] [CrossRef]
  102. Siegel, F.R. Factors influencing the precipitation of dolomitic carbonates. Geol. Surv. Kansas Bull. 1961, 152, 129–158. [Google Scholar]
  103. Brady, P.V.; Krumhansl, J.L.; Papenguth, H.W. Surface complexation clues to dolomite growth. Geochim. Cosmochim. Acta 1996, 60, 727–731. [Google Scholar] [CrossRef]
  104. Bontognali, T.R.R.; Vasconcelos, C.; Warthmann, R.J.; Bernasconi, S.M.; Dupraz, C.; Strohmenger, C.J.; McKenzie, J.A. Dolomite formation within microbial mats in the coastal sabkha of Abu Dhabi (United Arab Emirates). Sedimentology 2010, 57, 824–844. [Google Scholar] [CrossRef]
  105. Deng, S.C.; Dong, H.L.; Lv, G.; Jiang, H.C.; Yu, B.S.; Bishop, M.E. Microbial dolomite precipitation using sulfate reducing and halophilic bacteria: Results from Qinghai Lake, Tibetan Plateau, NW China. Chem. Geol. 2010, 278, 151–159. [Google Scholar] [CrossRef]
  106. Coniglio, M.; Frizzell, R.; Pratt, B.R. Reef-capping laminites in the Upper Silurian carbonate-to-evaporite transition, Michigan Basin, south-western Ontario. Sedimentology 2004, 51, 653–668. [Google Scholar] [CrossRef]
  107. Becker, F.; Bechstadt, T. Sequence stratigraphy of a carbonate-evaporite succession (Zechstein 1, Hessian Basin, Germany). Sedimentology 2006, 53, 1083–1120. [Google Scholar] [CrossRef]
  108. Liu, H.; Tan, X.C.; Li, Y.H.; Cao, J.; Luo, B. Occurrence and conceptual sedimentary model of Cambrian gypsum-bearing evaporites in the Sichuan Basin, SW China. Geosci. Front. 2018, 9, 1179–1191. [Google Scholar] [CrossRef]
  109. Sorento, T.; Olaussen, S.; Stemmerik, L. Controls on deposition of shallow marine carbonates and evaporites—Lower Permian Gipshuken Formation, central Spitsbergen, Arctic Norway. Sedimentology 2019, 67, 207–238. [Google Scholar] [CrossRef]
  110. Quijada, I.E.; Benito, M.I.; Suarez-Gonzalez, P.; Rodríguez-Martínez, M.; Campos-Soto, S. Challenges to carbonate-evaporite peritidal facies models and cycles; insights from Lower Cretaceous stromatolite-bearing deposits (Oncala Group, N Spain). Sediment. Geol. 2020, 408, 105752. [Google Scholar] [CrossRef]
  111. Corzo, A.; Luzon, A.; Mayayo, M.J.; van Bergeijk, S.A.; Mata, P.; García de Lomas, J. Carbonate Mineralogy Along a Biogeochemical Gradient in Recent Lacustrine Sediments of Gallocanta Lake (Spain). Geomicrobiol. J. 2005, 22, 283–298. [Google Scholar] [CrossRef]
  112. Brennan, S.T.; Lowenstein, T.K.; Cendón, D.I. The major-ion composition of Cenozoic seawater: The past 36 million years from fluid inclusions in marine halite. Am. J. Sci. 2013, 313, 713–775. [Google Scholar] [CrossRef]
  113. Wang, Y.; He, J.T.; Liu, C.C.; Chong, W.H.; Chen, H.Y. Thermodynamics versus Kinetics in Nanosynthesis. Angew. Chem. Int. Ed. 2015, 54, 2022–2051. [Google Scholar] [CrossRef] [PubMed]
  114. De Yoreo, J. Crystal nucleation more than one pathway. Nat Mater. 2013, 12, 284–285. [Google Scholar] [CrossRef] [PubMed]
  115. Gebauer, D.; Voelkel, A.; Coelfen, H. Stable prenucleation calcium carbonate clusters. Science 2008, 322, 1819–1822. [Google Scholar] [CrossRef]
  116. Meldrum, F.C.; Colfen, H. Controlling mineral morphologies and structures in biological and synthetic systems. Chem. Rev. 2008, 108, 4332–4432. [Google Scholar] [CrossRef]
  117. Meldrum, F.C.; Sear, R.P. Now you see them. Science 2008, 322, 1802–1803. [Google Scholar] [CrossRef]
  118. Raiteri, P.; Gale, J.D. Water is the key to nonclassical nucleation of amorphous calcium carbonate. J. Am. Chem. Soc. 2010, 132, 17623–17634. [Google Scholar] [CrossRef]
  119. Gebauer, D.; Coelfen, H. Prenucleation clusters and non-classical nucleation. Nano Today 2011, 6, 564–584. [Google Scholar] [CrossRef]
  120. Demichelis, R.; Raiteri, P.; Gale, J.D.; Quigley, D.; Gebauer, D. Stable prenucleation mineral clusters are liquidlike ionic polymers. Nat. Commun. 2011, 2, 590. [Google Scholar] [CrossRef] [PubMed]
  121. Radha, A.V.; Navrotsky, A. Thermodynamics of Carbonates. Rev. Mineral. Geochem. 2013, 77, 73–121. [Google Scholar] [CrossRef]
  122. Karthika, S.; Radhakrishnan, T.K.; Kalaichelvi, P. A Review of Classical and Nonclassical Nucleation Theories. Cryst. Growth Des. 2016, 16, 6663–6681. [Google Scholar] [CrossRef]
  123. Lutsko, J.F. How crystals form: A theory of nucleation pathways. Sci. Adv. 2019, 5, eaav7399. [Google Scholar] [CrossRef] [PubMed]
  124. Steefel, C.I.; Van Cappellen, P. A new kinetic approach to modeling water-rock interaction: The role of nucleation, precursors, and Ostwald ripening. Geochim. Cosmochim. Acta 1990, 54, 2657–2677. [Google Scholar] [CrossRef]
  125. Van Driessche, A.E.; Kellermeier, M.; Benning, L.G.; Gebauer, D. New Perspectives on Mineral Nucleation and Growth: From Solution Precursors to Solid Materials; Springer: Berlin/Heidelberg, Germany, 2016; p. 13. [Google Scholar]
  126. Nordeng, S.H.; Sibley, D.F. Dolomite stoichiometry and Ostwald’s step rule. Geochim. Cosmochim. Acta 1994, 58, 191–196. [Google Scholar] [CrossRef]
  127. Graf, D.L.; Goldsmith, J.R. Some hydrothermal syntheses of dolomite and protodolomite. J. Geol. 1956, 64, 173–186. [Google Scholar] [CrossRef]
  128. Goldsmith, J.R.; Graf, D.L. Structural and compositional variations in some natural dolomites. J. Geol. 1958, 66, 678–693. [Google Scholar] [CrossRef]
  129. Katz, A.; Matthews, A. The dolomitization of CaCO3. Geochim. Cosmochim. Acta 1977, 41, 297–308. [Google Scholar] [CrossRef]
  130. Sibley, D.F.; Bartlett, T.R.; Rodriguez-Clemente, R.; Tardy, Y. Nucleation as a rate limiting step in dolomitization. In Geochemistry and Mineral Formation in the Earth Surface; Rodriguez-Clemente, R., Tardy, Y., Eds.; Consejo Superior de Investigaciones Científicas: Madrid, Spain, 1987; pp. 733–741. [Google Scholar]
  131. Sibley, D.F.; Dedoes, R.E.; Bartlett, T.R. Kinetics of dolomitization. Geology 1987, 15, 1112–1114. [Google Scholar] [CrossRef]
  132. Usdowski, E. Synthesis of dolomite and geochemical implications. In Dolomites: A Volume in Honour of Dolomieu; Tucker, M.E., Zenger, D., Eds.; International Association of Sedimentologists Special Publication: Algiers, Algeria, 1994; Volume 21, pp. 345–360. [Google Scholar]
  133. Sibley, D.F. Unstable to stable transformations during dolomitization. J. Geol. 1990, 98, 739–748. [Google Scholar] [CrossRef]
  134. Sibley, D.F.; Nordeng, S.H.; Barkowski, M.L. Dolomitization kinetics in hydrothermal bombs and natural settings. J. Sediment. Res. 1994, 64, 630–637. [Google Scholar] [CrossRef]
  135. Kessels, L.A.; Sibley, D.F.; Nordeng, S.H. Nanotopography of synthetic and natural dolomite crystals. Sedimentology 2000, 47, 173–186. [Google Scholar] [CrossRef]
  136. Kaczmarek, S.E.; Sibley, D.F. A comparison of nanometer-scale growth and dissolution features on natural and synthetic dolomite crystals: Implications for the origin of dolomite. J. Sediment. Res. 2007, 77, 424–432. [Google Scholar] [CrossRef]
  137. Kaczmarek, S.E.; Sibley, D.F. On the evolution of dolomite stoichiometry and cation order during high-temperature synthesis experiments: An alternative model for the geochemical evolution of natural dolomites. Sediment. Geol. 2011, 240, 30–40. [Google Scholar] [CrossRef]
  138. Kaczmarek, S.E.; Sibley, D.F. Direct physical evidence of dolomite recrystallization. Sedimentology 2014, 61, 1862–1882. [Google Scholar] [CrossRef]
  139. Montes-Hernandez, G.; Findling, N.; Renard, F.; Auzende, A.L. Precipitation of ordered dolomite via simultaneous dissolution of calcite and magnesite: New experimental insights into an old precipitation enigma. Cryst. Growth Des. 2014, 14, 671–677. [Google Scholar] [CrossRef]
  140. Montes-Hernandez, G.; Findling, N.; Renard, F. Dissolution-precipitation reactions controlling fast formation of dolomite under hydrothermal conditions. Appl. Geochem. 2016, 73, 169–177. [Google Scholar] [CrossRef]
  141. Rodriguez-Blanco, J.D.; Shaw, S.; Benning, L.G. A route for the direct crystallization of dolomite. Am. Mineral. 2015, 100, 1172–1181. [Google Scholar] [CrossRef]
  142. Kaczmarek, S.E.; Thornton, B.P. The effect of temperature on stoichiometry, cation ordering, and reaction rate in high-temperature dolomitization experiments. Chem. Geol. 2017, 468, 32–41. [Google Scholar] [CrossRef]
  143. Kell-Duivestein, I.J.; Baldermann, A.; Mavromatis, V.; Dietzel, M. Controls of temperature, alkalinity and calcium carbonate reactant on the evolution of dolomite and magnesite stoichiometry and dolomite cation ordering degree—An experimental approach. Chem. Geol. 2019, 529, 119292. [Google Scholar] [CrossRef]
  144. Arvidson, R.S.; Mackenzie, F.T. The dolomite problem: Control of precipitation kinetics by temperature and saturation state. Am. J. Sci. 1999, 299, 257–288. [Google Scholar] [CrossRef]
  145. Malone, M.J.; Baker, P.A.; Burns, S.J. Recrystallization of dolomite: An experimental study from 50–200 °C. Geochim. Cosmochim. Acta 1996, 60, 2189–2207. [Google Scholar] [CrossRef]
  146. Kelleher, I.J.; Redfern, S.A.T. Hydrous calcium magnesium carbonate, a possible precursor to the formation of sedimentary dolomite. Mol. Simul. 2002, 28, 557–572. [Google Scholar] [CrossRef]
  147. Morse, J.W.; Casey, W.H. Ostwald processes and mineral paragenesis in sediments. Am. J. Sci. 1988, 288, 537–560. [Google Scholar] [CrossRef]
  148. Banerjee, A. Estimation of dolomite formation; dolomite precipitation and dolomitization. J. Geo. Soc. India 2016, 87, 561–572. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of calcite (a) and dolomite (b) crystal structures. The diagram is reproduced from Gregg et al. [38].
Figure 1. Schematic diagram of calcite (a) and dolomite (b) crystal structures. The diagram is reproduced from Gregg et al. [38].
Minerals 13 01479 g001
Figure 2. XRD patterns of ordered dolomite and very high-magnesium calcite (protodolomite). (a) (101), (015) and (021) ‘ordering’ reflections of ordered dolomite in XRD pattern; (b) XRD patterns of high-magnesium calcite (protodolomite). The diagrams are reproduced from Cai et al. [43] and Zheng et al. [5], respectively.
Figure 2. XRD patterns of ordered dolomite and very high-magnesium calcite (protodolomite). (a) (101), (015) and (021) ‘ordering’ reflections of ordered dolomite in XRD pattern; (b) XRD patterns of high-magnesium calcite (protodolomite). The diagrams are reproduced from Cai et al. [43] and Zheng et al. [5], respectively.
Minerals 13 01479 g002
Figure 3. Comparison of X–ray diffraction patterns of ordered dolomite and microbial dolomite [57,59,60,61,62,64,66]. The diagram is modified after Zhao et al. [58].
Figure 3. Comparison of X–ray diffraction patterns of ordered dolomite and microbial dolomite [57,59,60,61,62,64,66]. The diagram is modified after Zhao et al. [58].
Minerals 13 01479 g003
Figure 4. Distribution area of global surface water in dolomitization and dedolomitization at the CDGEP model. “α[Species]” represents the activity of the aqueous species. See references in Wang [24] for seawater data. The diagram is reproduced from Wang [24].
Figure 4. Distribution area of global surface water in dolomitization and dedolomitization at the CDGEP model. “α[Species]” represents the activity of the aqueous species. See references in Wang [24] for seawater data. The diagram is reproduced from Wang [24].
Minerals 13 01479 g004
Figure 5. Schematic diagram illustrating the catalytic role of negatively charged clay minerals in protodolomite formation: (a) diagrammatic crystal structure of 2:1 layer clay minerals (e.g., illite and montmorillonite); (b) the adsorption and dewatering of Mg and Ca ions by surface−bound hydroxyl groups; and (c) the formation of Mg(Ca)−hydroxyl complexes favoring the carbonation reaction. After Liu et al. [79].
Figure 5. Schematic diagram illustrating the catalytic role of negatively charged clay minerals in protodolomite formation: (a) diagrammatic crystal structure of 2:1 layer clay minerals (e.g., illite and montmorillonite); (b) the adsorption and dewatering of Mg and Ca ions by surface−bound hydroxyl groups; and (c) the formation of Mg(Ca)−hydroxyl complexes favoring the carbonation reaction. After Liu et al. [79].
Minerals 13 01479 g005
Figure 6. A schematic diagram showing the effect of sulfate on the formation of dolomite. (a) S O 4 2 binds Mg2+ to form tight CIP under hydrothermal conditions. TSR can remove sulfate to free Mg2+ and generate CO2 ( C O 3 2 / H C O 3 ), favoring the formation of hydrothermal dolomite; (b) S O 4 2 and Mg2+ mainly occur as hydrated ions and weakly associated ion pairs at surface temperatures. After Wang et al. [33].
Figure 6. A schematic diagram showing the effect of sulfate on the formation of dolomite. (a) S O 4 2 binds Mg2+ to form tight CIP under hydrothermal conditions. TSR can remove sulfate to free Mg2+ and generate CO2 ( C O 3 2 / H C O 3 ), favoring the formation of hydrothermal dolomite; (b) S O 4 2 and Mg2+ mainly occur as hydrated ions and weakly associated ion pairs at surface temperatures. After Wang et al. [33].
Minerals 13 01479 g006
Figure 7. Schematic diagram of carbonate nucleation by (a) classical nucleation theory through the addition of ions to a single cluster and (b) an alternative mechanism involving the aggregation of stable, amorphous, precritical clusters. After Meldrum and Sear [117].
Figure 7. Schematic diagram of carbonate nucleation by (a) classical nucleation theory through the addition of ions to a single cluster and (b) an alternative mechanism involving the aggregation of stable, amorphous, precritical clusters. After Meldrum and Sear [117].
Minerals 13 01479 g007
Figure 8. Micrographs of different minerals in synthesis experiments of dolomite. (a) Mg-calcite synthesized at ambient temperature; (b) protodolomite and first dolomite particles precipitated after 1 h of reaction, including a heating period from 20 to 300 °C; (c) ordered dolomite obtained at 300 °C for 48 h of reaction; (d) protodolomite; (e) details of the protodolomite in image (d); (f) dolomite; (g) details of dolomite crystallite subunit sizes in image (f). Diagrams (ac) are reproduced from Montes-Hernandez et al. [140]. Diagrams (eg) are reproduced from Rodriguez-Blanco et al. [141].
Figure 8. Micrographs of different minerals in synthesis experiments of dolomite. (a) Mg-calcite synthesized at ambient temperature; (b) protodolomite and first dolomite particles precipitated after 1 h of reaction, including a heating period from 20 to 300 °C; (c) ordered dolomite obtained at 300 °C for 48 h of reaction; (d) protodolomite; (e) details of the protodolomite in image (d); (f) dolomite; (g) details of dolomite crystallite subunit sizes in image (f). Diagrams (ac) are reproduced from Montes-Hernandez et al. [140]. Diagrams (eg) are reproduced from Rodriguez-Blanco et al. [141].
Minerals 13 01479 g008
Table 1. Thermodynamic parameters of dolomitization reactants and products in standard state (25 °C, 1 bar) as reported in the literature.
Table 1. Thermodynamic parameters of dolomitization reactants and products in standard state (25 °C, 1 bar) as reported in the literature.
Solid Phasef
(kJ·mol−1)
f
(kJ·mol−1)

(J·mol−1·K−1)
Reference
Dolomite−2162.354−2325.248154.890Berman [25]
−2161.7 ± 1.1−2324.5 ± 1.1 Hemingway and Robie [26]
−2161.51−2324.56156.0Holland and Powell [28]
−2147.82 ± 2.2 Rock et al. [29]
−2163.576−2329.86115.2Values given in Chen et al. [30]
−2160.9 ± 2−2323.1 ± 2156.9 ± 2Bénézeth et al. [19]
Calcite−1128.295−1206.81991.725Berman [25]
−1128.81−1207.5492.50Holland and Powell [28]
−1128.76−1206.8792.9Values given in Chen et al. [30]
Aragonite−1128.03−1207.6589.5Holland and Powell [28]
Gypsum−1797.2−2022.2193.8Nordstrom [31]
ionsf
(kJ·mol−1)
f
(kJ·mol−1)

(J·mol−1·K−1)
Reference
Ca2+−553.16−543.45−56.52Shock et al. [27]
−553.04−542.96−56.43Values given in Chen et al. [30]
Mg2+−454.29−466.27−138.16Shock et al. [27]
C O 3 2 −528.34−675.69−50.03Shock et al. [27]
H C O 3 −587.33−690.3998.49Shock et al. [27]
S O 4 2 −742.628 Nordstrom [31]
−744.96−910.2118.84Shock et al. [27]
H+000Shock et al. [27]
H2O(l)−237.141−285.8369.95Nordstrom [31]
Note: The ion data are the standard thermodynamic data in aqueous solution at 298.15 K; ∆fG°, standard Gibbs free energy of formation; ∆fH°, standard enthalpy of formation; S°, standard entropy
Table 2. The change in entropy, change in enthalpy and Gibbs free energy change for the dolomitization reactions.
Table 2. The change in entropy, change in enthalpy and Gibbs free energy change for the dolomitization reactions.
ReactionsrG1° = ∆fproducts − ∆freactants
(kJ·mol−1)
∆H
(kJ·mol−1)
∆S
(J·mol−1·K−1)
References to the Original Data
2CaCO3 + Mg2+ = CaMg(CO3)2 + Ca2+−2.1514.853.54Bénézeth et al. [19]; Shock et al. [27];
Holland and Powell [28]; Nordstrom [31]
CaCO3 + Mg2+ = CaMg(CO3)2−49.4626.4252.59
1.89CaCO3 + Mg2+ = CaMg(CO3)2 + 0.89Ca2+−66.94−58.0475.6
1.75CaCO3 + Mg2+ = CaMg(CO3)2 + 0.75Ca2+−146.06−151.2290.8
3 CaCO 3 ( s ) + M g 2 + + S O 4 2 + H+ + 2H2O = CaSO4· 2 H 2 O ( s ) + CaMg ( CO 3 ) 2 ( s ) + Ca 2 + + H C O 3 −13.808−140.62101.08
Table 3. Sulfate concentration where dolomite is still precipitated today vs. seawater sulfate concentration.
Table 3. Sulfate concentration where dolomite is still precipitated today vs. seawater sulfate concentration.
Field Sites SO 4 2 (mM)Reference
Early Holocene dolomite in Lake Sayram, Central Asia16.88Cheng et al. [8]
Hypersaline dolomitic lakes in the Coorong Region, S. Australia100.7–589.5Wright and Wacey [61]
Lake Jibuhulangtu Nuur, Inner Mongolia, China117.5Liu et al. [80]
Hypersaline coastal lagoon, Lagoa Vermelha, Brazil41–60Warthmann et al. [86]
Hypersaline coastal lagoon, Lagoa Vermelha, Brazil50Van Lith et al. [87]
Hypersaline coastal lagoon, Brejo do Espinho, Brazil69Van Lith et al. [87]
Pore water of slightly saline Qinghai Lake, NW China17.9Deng et al. [105]
Seawater~28.125Corzo et al. [111]
29Brennan et al. [112]
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

Chen, C.; Zhong, H.; Wang, X.; Ning, M.; Wang, X.; Ge, Y.; Wang, H.; Tang, R.; Hou, M. Thermodynamic and Kinetic Studies of Dolomite Formation: A Review. Minerals 2023, 13, 1479. https://doi.org/10.3390/min13121479

AMA Style

Chen C, Zhong H, Wang X, Ning M, Wang X, Ge Y, Wang H, Tang R, Hou M. Thermodynamic and Kinetic Studies of Dolomite Formation: A Review. Minerals. 2023; 13(12):1479. https://doi.org/10.3390/min13121479

Chicago/Turabian Style

Chen, Chao, Hanting Zhong, Xinyu Wang, Meng Ning, Xia Wang, Yuzhu Ge, Han Wang, Ruifeng Tang, and Mingcai Hou. 2023. "Thermodynamic and Kinetic Studies of Dolomite Formation: A Review" Minerals 13, no. 12: 1479. https://doi.org/10.3390/min13121479

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