Evaluating the Effects of Sodium, Potassium, Magnesium ...

81
Western Michigan University Western Michigan University ScholarWorks at WMU ScholarWorks at WMU Master's Theses Graduate College 12-2019 Evaluating the Effects of Sodium, Potassium, Magnesium, and Evaluating the Effects of Sodium, Potassium, Magnesium, and Calcium Concentrations on Dolomite Stoichiometry, Cation Calcium Concentrations on Dolomite Stoichiometry, Cation Ordering, and Reaction Rate Ordering, and Reaction Rate Hanna F. Cohen Follow this and additional works at: https://scholarworks.wmich.edu/masters_theses Part of the Geology Commons Recommended Citation Recommended Citation Cohen, Hanna F., "Evaluating the Effects of Sodium, Potassium, Magnesium, and Calcium Concentrations on Dolomite Stoichiometry, Cation Ordering, and Reaction Rate" (2019). Master's Theses. 5096. https://scholarworks.wmich.edu/masters_theses/5096 This Masters Thesis-Open Access is brought to you for free and open access by the Graduate College at ScholarWorks at WMU. It has been accepted for inclusion in Master's Theses by an authorized administrator of ScholarWorks at WMU. For more information, please contact [email protected].

Transcript of Evaluating the Effects of Sodium, Potassium, Magnesium ...

Western Michigan University Western Michigan University

ScholarWorks at WMU ScholarWorks at WMU

Master's Theses Graduate College

12-2019

Evaluating the Effects of Sodium, Potassium, Magnesium, and Evaluating the Effects of Sodium, Potassium, Magnesium, and

Calcium Concentrations on Dolomite Stoichiometry, Cation Calcium Concentrations on Dolomite Stoichiometry, Cation

Ordering, and Reaction Rate Ordering, and Reaction Rate

Hanna F. Cohen

Follow this and additional works at: https://scholarworks.wmich.edu/masters_theses

Part of the Geology Commons

Recommended Citation Recommended Citation Cohen, Hanna F., "Evaluating the Effects of Sodium, Potassium, Magnesium, and Calcium Concentrations on Dolomite Stoichiometry, Cation Ordering, and Reaction Rate" (2019). Master's Theses. 5096. https://scholarworks.wmich.edu/masters_theses/5096

This Masters Thesis-Open Access is brought to you for free and open access by the Graduate College at ScholarWorks at WMU. It has been accepted for inclusion in Master's Theses by an authorized administrator of ScholarWorks at WMU. For more information, please contact [email protected].

ii

EVALUATING THE EFFECTS OF SODIUM, POTASSIUM, MAGNESIUM, AND CALCIUM CONCENTRATIONS ON DOLOMITE STOICHIOMETRY, CATION

ORDERING, AND REACTION RATE

by

Hanna F. Cohen

A thesis submitted to the Graduate College in partial fulfillment of the requirements

for the degree of Master of Science Geosciences

Western Michigan University December 2019

Thesis Committee: Stephen E. Kaczmarek, Ph.D., Chair Peter Voice, Ph.D.

Andrew Caruthers, Ph.D.

iii

Copyright by Hanna F. Cohen

2019

i

ACKNOWLEDGMENTS

I would like to begin by thanking the Western Michigan University Department of

Geological and Environmental Sciences for providing funding for this project through a teaching

assistantship and the Western Michigan University College of Arts and Sciences for providing

funding through a summer research assistantship. I would also like to thank Dr. Kaczmarek for

his generous support through providing laboratory materials and space for all experiments and

analyses as well as his advising and writing support. I would also like to thank the Society for

Sedimentary Geology (SEPM), Geological Society of America – North Central Section, and

Geological Society of America – Mineralogy, Geochemistry, Petrology, and Volcanology

Division for travel funding to present this research at the annual Geological Society of America

conference in 2017 in Denver, CO and in 2018 in Seattle, WA. Finally, I would like to thank the

members of the Western Michigan University Carbonate Petrology and Characterization Lab for

ongoing collaboration and feedback throughout this project with additional gratitude to Brooks

Ryan for SEM training and feedback on my writing, and to Cameron Manche for XRD support.

Hanna F. Cohen

ii

EVALUATING THE EFFECTS OF SODIUM, POTASSIUM, MAGNESIUM, AND CALCIUM CONCENTRATIONS ON DOLOMITE STOICHIOMETRY, CATION

ORDERING, AND REACTION RATE

Hanna F. Cohen, M.S.

Western Michigan University, 2019

Numerous environmental factors affect dolomitization. Shallow peritidal and restricted

marine environments, for example, are often associated with more abundant and more

stoichiometric dolomite than deeper marine environments. Higher fluid Mg/Ca ratios resulting

from gypsum precipitation are often invoked to explain this observation, even when evidence of

evaporites is absent. In this study, high-temperature dolomitization experiments show that the

concentrations of major cation concentrations (Na, K, Mg, and Ca) impact dolomite

stoichiometry and reaction rate. Nearly 200 batch dolomitization experiments were run whereby

100 mg of natural aragonite ooids were dolomitized at 215°C in ionic solutions. Fluid [NaCl]

and [KCl] correlate positively with the stoichiometry of the initial protodolomite product (43–48

mol% MgCO3), but negatively with reaction rate. In contrast, the [Mg] and [Ca] of the

dolomitizing fluid correlate positively with both reaction rate and protodolomite stoichiometry

(41–45 mol% MgCO3). The rate at which cation ordering develops is unaffected by [NaCl],

[KCl], [Mg], or [Ca] in the dolomitizing fluid. These findings provide the basis for an alternative

explanation for the observed relationship between restricted, peri-tidal marine carbonate facies

and higher dolomite abundance and stoichiometry without the need to invoke precipitation of

calcium-bearing evaporites. These observations add to our understanding of the fundamental

controls on dolomite stoichiometry and reaction rate, and can help further constrain geological

interpretations based on dolomite stoichiometry.

ii

TABLE OF CONTENTS

ACKNOWLEDGEMENTS

LIST OF TABLES

LIST OF FIGURES

LIST OF EQUATIONS

CHAPTER

I. INTRODUCTION…………………………………………………………………….1

Dolomite ……………………………………………………………………..……..1

Research Question …………………………………………………………..........4

II. LITERATURE REVIEW …………………………………………………………….6

Fluid Chemistry ……………………………………………………………..........6

High Temperature Experimental Work ……………………………………..........8

III. METHODS ……………………………………………………………………...…..11

Experimental Procedure ……………………………………………….………..11

Analytical Techniques …………………………………………………………..15

IV. RESULTS …………………………………………………………………………....17

Replacement Reaction ……………………………………………….………….17

Protodolomite Reaction Rate ………………………………………………........20

Protodolomite Stoichiometry …………………………………………….……...22

Dolomite Cation Ordering …………………………………………………........24

V. DISCUSSION AND CONCLUSIONS ……………………………………………..26

Comparison to Previous Work ……………………………………………….…26

Reaction Curve ……………………………………………………….…26

iii

ii

v

vi

vii

iii

Table of Contents–Continued

CHAPTER

Rate of Dolomitization (Protodolomite) ………………………………...27

Ooid Size ……………………………………………………..…..……...28

Calcite Dolomitization………………………………………….………..28

Mechanisms for Salt Influence ………………………………………………….29

Implications ……………………………………………………………………...32

Dolomite as a Proxy for Understanding Fluid Chemistry ........................32

Global Correlations ……………………………………...............…........35

Conclusions ……………………………………………………………...………36

REFERENCES ………………………………………………………………………..………...38

APPENDIX

A. Dataset ………….……………………………...……………………..……….………... 45

iv

iv

LIST OF TABLES

1. Experimental values for each long-term solution.……………………………………12

v

v

LIST OF FIGURES

1. Model for the structure of well-ordered, stoichiometric dolomite ………...………….….3

2. Illustration of experiment set up.……………………………………………………..... 14

3. Illustration of experimental procedure ……………………….…………………………14

4. X-ray diffraction overlapping diffractograms for a pure aragonite sample and a pure dolomite sample …………………………………..……………………………….16

5. Percent product as a function of reaction time. ……………………………………........19

6. Stoichiometry as a function of percent Ca-Mg-carbonate product. ………………….....19

7. Percent product as a function of reaction time ………………………………………….20

8. Percent product as a function of sodium concentration in the dolomitizing fluid ..…….21

9. Percent product as a function of magnesium and calcium concentrations in the dolomitizing fluid …………..…………………………………………………………...22

10. Protodolomite stoichiometry as a function of sodium concentration in the dolomitizing fluid. ……………..………………………………………………………. 23

11. Protodolomite stoichiometry as a function of magnesium and calcium concentrations in the dolomitizing fluid ...………………..…………………………… 24

12. Percent protodolomite (no ordering reflections) as a function of reaction time …….....27

vi

vi

LIST OF EQUATIONS

1. Direct Precipitation Reaction…………………………………….………………………..2

2. Dissolution-Precipitation Reaction………………………………………………....……..2

3. Percent Product………………………………………………………………….….……16

4. Stoichiometry……………………………………………………………………….……16

5. Cation Ordering…………………………………………………………………….……16

vii

1

CHAPTER I INTRODUCTION

Dolomite

Dolomite, [CaMg(CO3)2], is a magnesium-calcium-carbonate mineral named in honor of

the French geologist Déodat Guy de Dolomieu, who was the first to document the occurrence of

a magnesium rich carbonate that did not readily react with hydrochloric acid as do limestones (de

Dolomieu, 1791). While dolomite is abundant in the rock record, modern (i.e. Holocene)

dolomite is extremely rare, and where present, is found only in minor quantities, in specific

settings. This observation, coupled with the failure to synthesize dolomite in the laboratory at

near surface conditions, is the crux of the “dolomite problem” (Van Tuyl, 1916; Sass and Bein,

1980; Land, 1980, 1985, 1998; Sun, 1994; Warren, 2000; Machel, 2004; Gregg et al., 2015;

Kaczmarek et al., 2017). Literature on the subject indicates wide disagreement about the

conditions, processes, and environments responsible for dolomite formation (Land, 1985;

Warren, 2000; Machel, 2004; Kaczmarek et al., 2017). Dolomite is the second most abundant

carbonate mineral in Earth’s crust and is economically significant as a hydrocarbon reservoir

rock (Warren, 2000; Machel, 2004). Due to the economic importance of dolomite and the lack of

modern naturally occurring dolomite, geologists routinely use proxies to infer the factors

controlling dolomitization in ancient rocks (Machel, 2004).

In nature, dolomite forms either as (i) an authigenic cement via direct precipitation from

an aqueous solution (Equation 1), or (ii) as a replacement product via a dissolution-precipitation

reaction whereby a CaCO3 precursor is replaced by dolomite in an Mg-bearing solution

(Equation 2). Dolomite cement is common in ancient carbonate rocks but is volumetrically minor

compared to platform-scale replacement dolomites observed throughout the Phanerozoic (Sun,

2

1994; Machel, 2004). The process whereby limestone is replaced by dolomite is referred to as

“dolomitization” (Equation 2).

Ca2+ + Mg2+ + 2CO32- à CaMg(CO3)2 (Equation 1)

Mg2+ + 2CaCO3 à CaMg(CO3)2 + Ca2+ (Equation 2)

Dolomite has a rhombohedral structure characterized by stacked layers of trigonal carbonate

anions separated by layers of calcium and magnesium cations (Figure 1; Lippmann, 1973;

Reeder, 1983; Gregg et al., 2015; Kaczmarek et al., 2017). The major elemental composition and

mineralogical structure of the mineral are described in terms of stoichiometry and cation

ordering, respectively – both of which can be measured quantitatively using x-ray diffraction

(Goldsmith and Graf, 1958; Royse et al., 1971; Lippmann, 1973; Lumsden, 1979; Reeder and

Sheppard, 1984; Zhang et al., 2010; Gregg et al., 2015; Kaczmarek et al., 2017). Dolomite

stoichiometry refers to the major cation composition in terms of the molar ratio of magnesium

relative to calcium in the crystal structure, commonly denoted as %MgCO3 (Lumsden, 1979;

Gregg et al., 2015). Dolomite stoichiometry is routinely measured using the position of the d-104

reflection in x-ray diffraction patterns (Lumsden, 1979; Zhang et al., 2010; Gregg et al., 2015).

Cation ordering is a measure of the degree to which the Mg and Ca cations are positioned in the

appropriate alternating cation layers within the dolomite lattice. Cation ordering is qualified by

the presence of the d-101, d-015, and d-021 ordering reflections in x-ray diffraction patterns

(Goldsmith and Graf, 1958; Gregg et al., 2015). An absence of these reflections indicates that a

Ca-Mg-carbonate mineral is not dolomite, but rather very high-magnesium calcite (Gregg et al.,

2015; Kaczmarek et al., 2017). Cation order is commonly semi-quantified using the intensity

ratio of the d-015 and d-110 x-ray diffraction reflections (Graf and Goldsmith, 1958; Gregg et

al., 2015). Well-ordered stoichiometric dolomite, for example, has equal molar proportions of Ca

3

and Mg segregated into alternating cation layers (Figure 1; Folk and Land, 1975; Warren, 2000;

Machel, 2004; Gregg et al., 2015; Kaczmarek et al., 2017).

Figure 1: Model for the structure of well-ordered, stoichiometric dolomite (Gregg et. al., 2015).

Despite stoichiometric dolomite being more stable at near-surface temperatures, most natural

dolomites have excess Ca (up to 12 mol%) in their lattice and thus some degree of cation

disorder (Goldsmith and Graf, 1958; Shinn et al., 1965; Illing et al., 1965; Lumsden, 1979;

Lumsden and Chimahusky, 1980; Sperber et al., 1984; Searl, 1994; Budd, 1997; Reeder, 2000;

Jones et al., 2001; Jones and Luth, 2002; Kaczmarek and Sibley, 2007; Turpin et al., 2012; Zhao

and Jones, 2012; Kaczmarek and Thornton, 2017; Ryan et al., 2019). Several factors have been

shown to affect the rate of dolomitization and dolomite stoichiometry. These include temperature

(Gaines, 1974; Usdowski, 1994; Thornton and Kaczmarek, 2015; Kaczmarek and Thornton,

2017), fluid Mg/Ca ratio (Land, 1967; Glover and Sippel, 1967; Sibley et al., 1987; Sibley, 1990;

Kaczmarek, 2005; Kaczmarek and Sibley, 2007, 2011), reactant mineralogy (Sibley et al., 1987;

Rose and Kaczmarek, 2019), reactant surface area (Sibley et al., 1987; Rose and Kaczmarek,

4

2019), organic matter (Zhang et al., 2012), sulfate reduction (Baker and Kastner, 1981; Morrow

and Rickets, 1988), pCO2 of the dolomitizing fluids (Sibley, 1990), and recrystallization

(Kaczmarek and Sibley, 2014; Ryan et al., 2019). Various case studies have presented evidence

suggesting that dolomite composition directly reflects the geochemistry of the local environment

of formation (Sass and Bein, 1988; Folk and Land, 1975; Lumsden and Chimahusky, 1980;

Kaczmarek and Sibley, 2011; Ren and Jones, 2017; Manche and Kaczmarek, 2019a).

Research Question

This study focuses on the question of how do the concentrations of common cations in

dolomitizing fluids, namely sodium, potassium, magnesium, and calcium, which all increase

during evaporation, influence dolomite stoichiometry and reaction rate. Based on observations

from natural environments, it is hypothesized that increasing salinity ([NaCl], [KCl], [Mg], and

[Ca]) of the fluid will increase the rate of dolomitization as well as the stoichiometry and cation

ordering of the resulting dolomite. High-temperature dolomitization experiments are commonly

used to understand the effects of various parameters on the dolomitization reaction (e.g., Glover

and Sippel, 1967; Liebermann, 1967; Land, 1967; Baker and Kastner, 1981; Sibley et al., 1987;

Sibley, 1990; Zempolich and Baker, 1993; Malone et al., 1996; Kaczmarek and Sibley, 2007,

2011, 2014; Zhang et al., 2012; Kaczmarek and Thornton, 2017). With high-temperature

experiments, individual factors influencing dolomitization can be controlled rather than inferred.

To date, only two studies have set out to directly evaluate how common environmental

conditions impact dolomite stoichiometry (Kaczmarek and Sibley, 2011; Kaczmarek and

Thornton, 2017). The results from the experiments presented in this thesis will provide a new

5

way to explain the higher abundance of stoichiometric dolomite in evaporitic depositional

settings.

6

CHAPTER II

LITERATURE REVIEW

Fluid Chemistry

In nature, stoichiometric dolomite is commonly associated with evaporitic depositional

environments (e.g., Goldsmith and Graf, 1958; Folk and Land, 1975; Sass and Bein, 1988).

Evaporitic settings also tend to contain more abundant dolomite than their subtidal counterparts

(Goldsmith and Graf, 1958; Folk and Land, 1975; Machel, 2004). Such observations are

routinely explained by invoking elevated fluid Mg/Ca ratios caused by sedimentary gypsum

precipitation in restricted peritidal settings (Carpenter, 1980; Machel, 2004). This explanation is

geologically reasonable as it fits with our understanding of the conditions that promote dolomite

formation. More specifically, gypsum (CaSO4 • 2H2O) precipitation should drive fluid Mg/Ca

ratios higher, which is more thermodynamically favorable for dolomite (Carpenter, 1980).

However, the main limitations of this model are such that (i) in addition to a Mg/Ca ratio

increase, other chemical changes take place during evaporation of marine waters, and (ii) many

restricted marine settings lack evidence of evaporite precipitation. Folk and Land (1975) studied

natural dolomites and inferred that with increasing concentrations of Mg, Ca, and Na in the

solution, a significant increase in the Mg/Ca ratio from the precipitation of evaporites would also

be required for well-ordered, stoichiometric dolomite to form. They hypothesized that the Mg/Ca

ratio in the fluid was a determining factor on the dolomitization reaction. Sass and Bein (1988)

studied natural Permian to Neogene dolomites from Israel and Texas (U.S.) with varying sodium

contents and concluded that the chemical and mineralogical makeup of natural dolomites directly

reflected the geochemistry of their depositional and diagenetic environments. They inferred

water chemistry (i.e. the degree of evaporation) from the [Na] in the dolomite and measured

7

stoichiometry using XRD. Despite very contrasting sodium concentrations in the dolomitizing

fluids, they showed that the dolomite stoichiometries from marine and gypsum saturated fluids

evenly spanned the same range of 44-50 %MgCO3. Sass and Bein (1988) also reported, however,

that dolomite associated with halite saturated brines was more stoichiometric (49-50 %MgCO3).

They interpreted, as did others previously, that the observed differences likely reflected different

Mg/Ca ratios in the dolomitizing fluids. More specifically, that more stoichiometric dolomite

was formed in restricted marine to evaporitic diagenetic settings because of higher fluid Mg/Ca

associated with the precipitation of gypsum (Goldsmith and Graf, 1958; Folk and Land, 1975;

Sass and Bein, 1988).

Past laboratory experiments focused on how fluid chemistry influences the rate of

dolomitization. Land (1967), for example, performed a series of high-temperature dolomitization

experiments and reported an increase in rate of dolomitization with increased concentrations of

magnesium and calcium. Stoichiometry and cation ordering data were not reported in this study,

however. In other laboratory experiments, Liebermann (1967) observed that the rate of

dolomitization of a CaCO3 precursor decreased as concentrations of NaCl, MgCl2, MgSO4,

CaSO4, and KCl in artificial seawater increased. Liebermann (1967) interpreted his observations

to indicate that the additives caused the solubility of dolomite to decrease and the solubility of

CaCO3 to increase. No data on stoichiometry or cation ordering were reported. Glover and Sippel

(1967) synthesized aragonite and high-magnesium calcite via direct precipitation from fluids in

high-temperature experiments using aqueous solutions containing [Mg], [Ca], and [Na] like

modern seawater and reported that a lower pH (i.e. 6–7 pH) might contribute to the higher total

carbonate concentration in the formation of primary dolomite. Glover and Sippel (1967) did not

disqualify a higher pH and thus a lower total carbonate concentration from forming primary

8

dolomite, but did recognize it would require significantly more time. Sibley et al. (1994) reported

no change in dolomite stoichiometry in experiments spiked with NaCl, although reaction rate

data were not reported. More recently, Martin et al. (2013) and Kaczmarek et al. (2015), reported

that increased concentrations of NaCl in the dolomitizing fluid correlated with a decreased rate

of dolomitization of a calcite precursor and an increase in the stoichiometry of the initial

products. Considering contradictory results in the literature, the objective of this study is to more

thoroughly and systematically evaluate the effects of [NaCl], [KCl], [Mg], and [Ca], major

chemical components in modern seawater, on dolomite stoichiometry and cation ordering.

High Temperature Experimental Work

In general, it is known that there are many geochemical parameters that affect dolomitization

including temperature, fluid Mg/Ca ratio, molarity of the fluid with respect to various cations

(Na, K, Mg, Ca), SO42-, microbial activity, and reactant composition and texture (Liebermann,

1967; Land, 1967; Lippmann, 1973; Gaines, 1974; Folk and Land, 1975; Baker and Kastner,

1981; Sibley et al., 1987; Machel, 2004; Kaczmarek and Sibley, 2007, 2011, 2014; Kaczmarek

and Thornton, 2017; Rose and Kaczmarek, 2019). In most diagenetic environments, these

geochemical parameters evolve concurrently, making it difficult to evaluate their individual

effects. Laboratory experiments, in contrast, provide an approach for constraining individual

variables and directly examining their effects. The impact of reaction temperature on the rate of

dolomitization, for example, has been studied extensively (e.g., Gaines, 1974; Usdowski, 1994;

Kaczmarek and Thornton, 2017), but only one such study has examined how temperature

impacts dolomite stoichiometry (Kaczmarek and Thornton, 2017). Similarly, the effect of fluid

Mg/Ca was studied by Land (1967), Sibley et al., (1987), and Kaczmarek and Sibley (2011).

9

These studies all showed that an increase in the Mg/Ca ratio results in an increase in reaction

rate. Only two of these studies examined dolomite stoichiometry (Sibley et al, 1987; Kaczmarek

and Sibley, 2011). Kaczmarek and Sibley (2011) evaluated the effects of Mg/Ca and they

observed an increase in rate of dolomitization with increased Mg/Ca. Sulfate concentrations have

also been studied in the laboratory and were determined by Baker and Kastner (1981) to inhibit

dolomitization, even at very small concentrations (less than 5%). As sulfate is reduced, alkalinity

increases, NH4+ is produced, and, in exchange, additional Mg2+ is released from solid silica and

dissolves into the dolomitizing fluid (Baker and Kastner, 1981; Morrow and Rickets, 1988).

The application of high-temperature experiments to understanding the factors affecting the

formation of natural near-surface dolomites is supported in literature by the observed similarities

in mineralogy, stoichiometry, and crystal morphology of both natural and synthetic dolomites

(Sibley et al., 1987; Zempolich and Baker, 1993; Kaczmarek and Sibley, 2014, 2017; Kaczmarek

and Thornton, 2017). The disadvantage of using experiments as an analogue to understanding

ancient dolomites is that high reaction temperatures speed up the reaction and may impact the

dolomitization reaction in a way that is not completely understood (Kaczmarek and Sibley, 2014;

Kaczmarek at al., 2017). The petrological similarities between experimental and natural

dolomites in literature imply that experimental dolomites can be used as an effective analogue

for natural sedimentary dolomites. The mimetic replacement of allochems in high-temperature

dolomitization experiments appears to be the same as the textural preservation in some naturally

dolomitized allochems (Sibley et al., 1987). Sibley et al. (1987) observed similarities in the

product dolomite, including rhombic morphology of dolomite formed in high temperature

experiments that matches natural dolomite. Sibley et al. (1897) also recognized the scarcity of

partially dolomitized rocks due to the reaction rate observed in high-temperature experiments

10

and an increase in stoichiometry in completely dolomitized limestone (Sibley et al., 1987).

Zempolich and Baker (1993) and Kaczmarek and Thornton (2017) observed textural similarities

in dolomitized ooids from nature and those produced in high-temperature experiments. The

nanotopography of the surface of natural and experimental product dolomite surfaces is also the

same (Kaczmarek and Sibley, 2007, 2014).

11

CHAPTER III

METHODS

Experimental Procedure

To evaluate the hypotheses, 189 high-temperature dolomitization (dissolution-

precipitation) experiments were carried out to assess the effects of Na, K, Ca, and Mg

concentrations on dolomite composition and reaction rate. The laboratory experiments followed

a similar approach used previously in Kaczmarek and Sibley (2007, 2011, 2014) and Kaczmarek

and Thornton (2017). Batch dolomitization experiments were conducted whereby aragonite

ooids were dolomitized in Mg-Ca-Cl fluids exhibiting a range of salinities. A stock electrolyte

solution was produced by dissolving 10 moles (2033 g ± 10 g) of magnesium dichloride

hexahydrate and 10 moles (1110 g ± 10 g) of calcium dichloride dehydrate in 10 liters of ultra-

pure H2O. The mixtures were mechanically stirred for approximately 10 minutes to ensure all

solutes were dissolved. The calculated molarity of the stock solution with respect to Mg2+ and

Ca2+ is 0.86 mol/L. The Mg2+/Ca2+ ratio of the stock solution is 1.0.

The stock solution was divided equally into five Erlenmeyer flasks and different

quantities of salt ([NaCl] or [KCl]) were dissolved in each flask to obtain experimental solutions.

The experimental solutions were created to mimic the [Na], [K], [Mg], and [Ca] concentrations

of common dolomitization fluids (e.g., seawater and hypersaline brines) (Table 1; Logan, 1987;

Jones and Xiao, 2005). Cation concentrations are based on previously published data from

surface waters in Lake MacLeod, an intracratonic evaporite basin in western Australia (Table 1;

Logan, 1987; Jones and Xiao, 2005). Lake MacLeod is a modern lake exhibiting continued

reflux due to continuous evaporation of seawater coupled with the periodic flooding of meteoric

water (Logan, 1987). The fluid chemistry of Lake MacLeod was previously used for reactive

12

transport modeling (RTM) (Jones and Xiao, 2005; Gabellone and Whitaker, 2016). Gabellone

and Whitaker (2016) used reactive transport modeling (coupled chemical and hydrological

numerical model) and hypothesized an expected increase in dolomitization rate with an increase

in concentration of total dissolved solids. Later work by Kaczmarek et al. (2016) showed that the

geochemical models used in RTM do not specifically consider [NaCl] and [KCl]. In this study, a

series of high-temperature experiments are used that directly measure the effect of [NaCl],

[KCl], [Mg], and [Ca] on the stoichiometry and rate of dolomitization.

Table 1: Experimental values for each long-term solution. Values based on those of Logan (1987).

Two sets of experiments – long-term and short-term – were used to determine the effects of

solution chemistry on dolomite composition and reaction rate. Long term experiments were

conducted to elucidate the effects of different fluid [NaCl] and [KCl] on long-term

dolomitization rate and how dolomite stoichiometry and cation ordering evolve during the multi-

stage reaction. 160 experiments lasting between 2 and 800 hours were conducted. In these

experiments, aragonite ooids were reacted in high-temperature (215°C) dolomitizing fluids that

were synthesized to mimic zero salt, seawater, and evaporated seawater (hypersaline) conditions

(Table 1). The products were analyzed and percent dolomite, stoichiometry, and cation ordering

[Mg]

[Ca]

[Na]

[K]

Stock Solution Modern Seawater Hypersaline KCl (NaCl)

0.86 mol/L

0.86 mol/L

N/A

0.484 mol/L

0.86 mol/L

0.86 mol/L

2.93 mol/L

N/A

0.86 mol/L

0.86 mol/L

0.485 mol/L

N/A

0.86 mol/L

0.86 mol/L

N/A

N/A

13

were measured to address the long-term effect of changing fluid [NaCl] and [KCl] on

dolomitization (Appendix A).

A series of short term (5 and 10 hour) dolomitization experiments were also conducted to

evaluate the sensitivity of stoichiometry with smaller intervals of concentrations of Na, K, Mg,

and Ca. The Na experiments were 5 hours in length while the Mg and Ca experiments were 10

hours in length. Five-hour [Na] experiments were intended to evaluate the stoichiometry and

reaction rate of the initial product. For the 5 hour experiments, varying amounts (0.0-3.5 g) of

NaCl were added in 0.5 g increments to 15 mL of stock solution. This included replicates of the

solutions created for the long-term experiments. For the short-term experiments evaluating [Mg]

and [Ca], concentrations of Mg and Ca were added to the stock solution up to a molarity of 1.5

mol/L. The experimental solutions used in these experiments had Mg/Ca ratio of 1. Short-term

experiments evaluating the effects of [Mg] and [Ca] were ten hours in duration to analyze the

initial product.

Naturally occurring modern aragonite ooids were used as the solid reactants in all

experiments (Appendix A). Ooids were harvested from the Ambergris Shoal, Turks and Caicos

Islands (Kaczmarek and Thornton, 2017). The ooids were sieved to obtain the 355-425µm size-

fraction to control for reactant surface area.

Each Teflon-lined stainless steel reaction vessel (bomb) was loaded with 100 mg of

aragonite ooids and 15 mL of the experimental solution (Figure 2). Reaction vessels were sealed,

immediately placed in a high precision Thermo Scientific Heratherm convection oven preheated

to 215°C. Individual vessels were removed periodically between 2-800 hours, and were force

cooled to room temperature with compressed air. Solids were obtained by filtering using a

vacuum flask. Solids were rinsed with ~500 mL of de-ionized water to remove any salt

14

precipitates. Experimental fluids were discarded. Solids were dried for a minimum of 30 minutes

in a vacuum desiccator prior to further analysis (Figure 3). These methods have been previously

described in Cohen and Kaczmarek (2017) and Cohen and Kaczmarek (2018).

Figure 2: Illustration of experiment set up. Teflon-lined stainless steel reaction vessel with a flanged closure loaded with experimental solution (blue) and aragonite ooids (white) Adapted from Parr Instrument Company (2019).

Figure 3: Illustration of experimental procedure. Procedure follows: (1) solution preparation, sieving of ooids, and loading of bombs, (2) placement of bombs in oven, (3) removal of samples to dry and desiccate, (4) powder products, and (5) analyze powdered products with x-ray diffractometer. Adapted from Thermo Fisher Scientific Inc. (2012), DWK Life Sciences LLC. (2017), Cole-Parmer Instrument Company, LLC (2019), and Parr Instrument Company (2019).

15

Analytical Techniques

The mineralogical composition of the solid products was analyzed using standard powder x-

ray diffraction techniques. X-ray diffraction (XRD) is an analytical tool routinely used to

determine the relative proportions of carbonate minerals in a sedimentary assemblage. It is also

routinely used to determine the stoichiometry and cation ordering of dolomites (Royse et al.,

1971; Lumsden, 1979; Reeder and Sheppard, 1984; Zhang et al., 2010; Gregg et al., 2015). To

homogenize the samples, solid products were powdered by hand using an agate mortar and pestle

and then mounted on a Boron-doped silicon P-type zero background diffraction plate. Each

sample was analyzed with a Bruker D2 Phaser diffractometer at Western Michigan University

using CuKα radiation (λ=1.54184 Å) at 300 Watts (30 kV, 10 mA). The diffractometer uses

Bragg-Brentano geometry with a fixed linear LYNXEYE 1D line detector. The samples were

scanned over the range of 20-55°2θ with a 0.008°2θ step size increment (4327 steps), at a speed

of 1.0 sec/step. Samples were rotated at 10 revolutions/minute to ensure that the entire sample

was analyzed. Peak positions were confirmed with an internal standard. The scan parameters

were modified from those used in Kaczmarek and Thornton (2017).

XRD patterns were analyzed with the Bruker EVA toolkit with the Crystallography Open

Database. Patterns were processed to remove k-beta, background, and smoothed using the

Bruker EVA toolkit and are consistent with conventional data processing techniques (Kaczmarek

and Thornton, 2017). These data were also used to determine the mineralogical composition (i.e.

percent Ca-Mg-carbonate products relative to aragonite reactants), dolomite stoichiometry, and

cation ordering based on the positions and intensities of specific reflections (Figure 4; Royse et.

al., 1971; Goldsmith and Graf, 1958; Lumsden, 1979). The percent product was determined

using the ratio of the [104] peak intensity to the sum of the intensities of the [104] and aragonite

16

[111] peaks (Equation 3). Dolomite stoichiometry was quantified based on the position of the

[104] peak using the empirical relationship (Equation 4) established by Lumsden (1979). Mg

ions are smaller than Ca ions and thus cause a reduction in the d-spacing of the layer in the

crystal structure and an increase in the 2-theta value for the dolomite [104] peak, per Bragg’s

Law (Lumsden, 1979). Cation ordering was determined by the intensity ratio of the dolomite

[015] ordering peak to the dolomite [110] peak (Equation 5), as detailed in Goldsmith and Graf

(1958).

Figure 4: X-ray diffraction overlapping diffractograms for a pure aragonite sample and a pure dolomite sample. Aragonite sample is in light blue and dolomite sample is in dark blue. Peaks used for determining percent dolomite, stoichiometry, and cation ordering labeled.

% dolomite = !"#$%(()*+),-*.)!"#$% ()*+),-*. 01"###(()*+),-*.)

∗ 100 (Equation 3)

mole% MgCO3 = 1 – ((333.33*(D {104} d-space))-911.99) (Equation 4)

cation ordering = {015} / {110} (Equation 5)

17

CHAPTER IV

RESULTS

To determine experimental error, 7 replicates were run for 18 hours in the stock solution.

Based on the data from these runs, the analytical uncertainty for dolomite stoichiometry and

percent product is ±0.538 mol% MgCO3 and ±2.38% product, respectively (Appendix A).

Replacement Reaction

Replacement reaction rate data from the long-term high-temperature dolomitization

experiments indicate an increased abundance of Ca-Mg-carbonate product with reaction time.

Data plotted in Figure 5 show a characteristic S-shaped reaction curve proceeding in a non-linear

fashion. Reaction data from all four experimental solutions (stock solution, NaCl seawater, NaCl

hypersaline, and KCl seawater) follow the same general pattern previously defined by

Kaczmarek and Sibley (2014), whereby the reaction proceeds through four distinct stages that

represent a progressive replacement of aragonite by a protodolomite, poorly ordered dolomite,

and well-ordered dolomite (Figure 5). The four distinct reaction stages previously defined by

Kaczmarek and Sibley (2014) will be referred to here as the induction stage, rapid replacement

stage, primary recrystallization stage, and secondary recrystallization stage (Figure 5). During

the induction stage, no products are detected by XRD. The induction stage in all three

experiments lasted between three and four hours (Appendix A). This means that in the

experiments which ended within 3 hours of reaction time, no products were detected. After 4

hours, however, Ca-Mg-carbonate products were detected by XRD in all experiments. The

stoichiometry of these initial Ca-Mg-carbonate phases ranges from 41.0 – 46.2 mole% MgCO3.

No cation ordering in these products was detected with XRD. Based on the dolomite like

18

composition, but lack of cation ordering, these phases can be referred to as either very high-

magnesium calcite, sensu Gregg et al. (2015), or protodolomite, sensu Graf and Goldsmith

(1956). Herein, the initial Ca-Mg-carbonate phases observed in these experiments will be

referred to as protodolomite because it can be demonstrated that they become ordered dolomite

given enough time. The rapid replacement stage is characterized by fast protodolomite formation

as evidenced by a steep slope in Figure 6. During the rapid replacement stage the products

maintain a relatively constant stoichiometry until ~95% product as shown by the nearly flat

slopes (m<0.02) and low correlation coefficients shown in the plot of dolomite stoichiometry vs.

time (Figure 6, Appendix A). The subsequent primary recrystallization stage, in contrast, is

characterized by a marked decrease in reaction rate as evidenced by the reduction in slope in

Figure 5. A rapid increase in stoichiometry to 50 mol% MgCO3 occurs during the primary

recrystallization stage, as indicated by the steep slopes in Figure 6. The secondary

recrystallization stage is characterized by dolomite products that are have become fully

stoichiometric but cation ordering that continues to increase with time to some maximum value

and then levels off (Appendix A). The end products from the longest experiments, which lasted

800 hours, were well-ordered and stoichiometric dolomite (Appendix A).

19

Figure 5: Percent product as a function of reaction time. Percent product refers to dolomite relative to aragonite. Data is consistent with previously published model (Kaczmarek and Sibley, 2014).

Figure 6: Stoichiometry as a function of percent Ca-Mg-carbonate product. Prior to recrystallization, stoichiometry is relatively constant. Data is consistent with previously published model (Kaczmarek and Thornton, 2017).

Aragonite Poorly Ordered DolomiteProtodolomite Well Ordered

Dolomite100

80

60

40

20

00 10 100 1000

Reaction time (hours)

% P

rodu

ct

Induction

Rapi

d Re

plac

emen

t

Primary and Secondary Recrystallization

Mg2+ + 2CaCO3 --> CaMg(CO3)2 + Ca2+

Hypersaline: m=51.67, R2 = 0.9217Modern Seawater: m=48.12, R2 = 0.9654Stock Solution: m=49.02, R2 = 0.9721

1. Induction: nucleation and slow growth2. Rapid Replacement: protodolomite growth3. Primary Recrystallization: dolomite stoichiometry and cation ordering increase4. Secondary Recrystallization: fully stoichiometric dolomite, cation order increases to maximum

52

50

48

46

44

42

40

% Product

MgC

O3 m

ol %

0 20 40 60 80 100

Stoichiometry vs. % Product Rapid Replacement Primary RecrystallizationHypersaline: y=0.01x+44.6 R2 = 0.119 y=0.85x-34.1 R2 = 0.657Modern Seawater: y=-0.0007x+43.8 R2 = 0.009 y=0.68x-17.2 R2 = 0.607Stock Solution: y=0.02x+42.3 R2 = 0.297 y=1.19x-67.2 R2 = 0.778

Rapid Replacement

Prim

ary

Rec

ryst

alliz

atio

n

20

Protodolomite Reaction Rate

Reaction rates during the rapid replacement stage were determined based on the percent

product per reaction time (Figure 5; Figure 7; Appendix A). The rapid replacement stage in these

experiments is defined as the end of the induction period until 60% protodolomite (Figures 5–7).

In long-term experiments, a general trend is observed of slightly faster protodolomite formation

in the stock solution than in the solutions with added salt (i.e. NaCl seawater, NaCl hypersaline,

and KCl). Experiments conducted in the stock solution display a reaction rate slope of 7.19

whereas the slopes of the reaction rate for the NaCl seawater, NaCl hypersaline, and KCl

experiments are 7.12, 6.96, and 6.80, respectively. Data from the individual experiments have

high correlation coefficients (R2 = 0.97 – 0.99), and the addition of NaCl and KCl decreases the

rate of dolomitization in comparison to the stock solution with no added salt (Figure 7).

Figure 7. Percent product as a function of reaction time. Data plotted from the four long-term experiments up to 60% protodolomite. The zero salt shows the fastest rate, but the slopes are very similar indicating only minor differences.

60

50

40

30

20

10

00 2 4 6 8 10 12 14 Time (Hrs)

% p

roto

dolo

mite

NaCl Hypersaline y=6.9x - 28.1 R2 = 0.99NaCl Seawater y=7.1x - 23.4 R2 = 0.99KCl y=6.8x - 23.5 R2 = 0.98Stock (No salt) y=7.2x - 20.9 R2 = 0.97

21

Results from the 5-hour Na experiments also show that reaction rate is inversely

proportional to the NaCl concentration in the solution (Figure 8). After 5 hours, 6.86%

protodolomite formed in the solution with 2.57 g NaCl and 15.66% protodolomite formed in the

solution with 0.25 g NaCl (Figure 8; Appendix A).

Figure 8: Percent product as a function of sodium concentration in the dolomitizing fluid. Percent product determined by x-ray diffraction analysis. Data plotted from short-term experiments evaluating [NaCl]. Regression line shows rate of dolomitization in the early stage of the reaction decreases with increasing concentrations of Na. [KCl] data has a similar trend. Blue points show results of 300-555 µm Iceland spar calcite experiments (Martin et al., 2013; Kaczmarek et al., 2016) Black dots show results of 355-425 µm aragonite ooid experiments (this study).

Stoichiometry (5 h)

MgC

O3 m

ol %

NaCl mass in experimental solution (g/15mL)

49

48

47

46

45

44

43

42 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5

300-355 μm: R2 = 0.6900355-425 μm: R2 = 0.5159

30%

25%

20%

15%

10%

5%

0%

NaCl mass in experimental solution (g/15mL)

% P

rodu

ct

0 0.5 1.0 1.5 2.0 2.5 3.0 3.5

Rate of Dolomitization (5 h)300-355 μm : R2 = 0.7800355-425 μm: R2 = 0.8055

22

Results from the 10-hour Mg and Ca experiments show that as concentrations of Mg and

Ca increase, the rate of dolomitization increases (Figure 9) below concentrations of 1.0 mol/L of

Mg and Ca. After 10 hours, 9.52% initial product formed in the 0.05 M solution and 56.41%

initial product formed in the 0.87 M solution (R2=0.91) (Figure 9; Appendix A).

Figure 9: Percent product as a function of magnesium and calcium concentrations in the dolomitizing fluid. Percent product determined by x-ray diffraction analysis. Data plotted from short-term experiments evaluating [Mg] and [Ca] demonstrating an increased rate of dolomitization with increasing concentrations. The shaded portion represents values that exceed those found in natural environments.

Protodolomite Stoichiometry

In the long-term experiments the stoichiometry of the initial Ca-Mg-carbonate products,

i.e. protodolomite, (Figure 6) is higher with increased concentrations of salt in the fluid. The

average stoichiometry of the protodolomite produced in the hypersaline solution is 44.93 mol%

MgCO3. In contrast, the average stoichiometry of the protodolomite produced in the seawater

and stock solutions is lower at 43.69 mol% MgCO3 and 43.19 mol% MgCO3, respectively

(Appendix A). The average stoichiometry of the protodolomite produced in the KCl solution is

23

43.55 mol% MgCO3. The addition of NaCl and KCl both increase the average stoichiometry of

the initial product in comparison to the stock solution with no added salt.

Data from the short-term experiments show a similar relationship between salt

concentration and stoichiometry (Figure 10). The 5-hour Na experiments show that the

stoichiometry of the initial products positively correlates with the amount of NaCl added to the

solution (0-2.57 g per 15 ml). The protodolomite products of the 5-hour experiments have

stoichiometry values that range from 43.80 – 45.89 mol% MgCO3 over the range of NaCl

examined and have a corresponding R2 value of 0.52 (Figure 10).

Figure 10: Protodolomite stoichiometry as a function of sodium concentration in the dolomitizing fluid. Fluid chemistry plotted from short-term experiments showing an increase in stoichiometry with increasing [NaCl] and [KCl]. Only [NaCl] data is plotted. [KCl] data is consistent, but not show on the graph. NaCl concentration (g NaCl/15 mL stock solution) is plotted on the x-axis and stoichiometry is plotted on the y-axis. Solid black dots are from the experiment using 355-425 µm aragonite ooids. Solid blue dots are from experiments with an identical procedure but smaller ooids.

Stoichiometry (5 h)

MgC

O3 m

ol %

NaCl mass in experimental solution (g/15mL)

49

48

47

46

45

44

43

42 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5

300-355 μm: R2 = 0.6900355-425 μm: R2 = 0.5159

30%

25%

20%

15%

10%

5%

0%

NaCl mass in experimental solution (g/15mL)

% P

rodu

ct

0 0.5 1.0 1.5 2.0 2.5 3.0 3.5

Rate of Dolomitization (5 h)300-355 μm : R2 = 0.7800355-425 μm: R2 = 0.8055

24

Results from the 10-hour Mg and Ca experiments show that as concentrations of Mg and

Ca increase, dolomite stoichiometry increases below [Ca] and [Mg] of 1.0 mol/L, but increases

above 1.0 mol/L (Figure 11). Protodolomite stoichiometry in the 0.05 M [Ca] and [Mg] solution

is 41.19 mol% MgCO3 compared to 44.27 mol% MgCO3 in the 0.87 M [Mg] and [Ca] solution.

An R2 value of 0.87 is shown for the data below 1.0 mol/L (Figure 11; Appendix A).

Figure 11: Protodolomite stoichiometry as a function of magnesium and calcium concentrations in the dolomitizing fluid. Fluid chemistry from short-term experiments evaluating [Mg] and [Ca] demonstrating an increase in stoichiometry with increasing concentrations. The shaded portion represents values that exceed those found in natural environments.

Dolomite Cation Ordering

The 5-hour Na experiments and the 10-hour Mg and Ca experiments only produced

initial products and no cation ordering was observed. All solutions required at least 22 hours

before XRD evidence of cation ordering was observed (Appendix A). The long-term

experiments show that the cation ordering of the dolomite formed is directly influenced by the

25

concentration of salt in the fluid. The dolomite formed from the stock solution with zero salt

added reaches a maximum cation ordering of 0.34. The dolomite formed from the seawater

solution reaches a maximum cation ordering of 0.34. The dolomite formed from the hypersaline

solution reaches a maximum cation ordering of 0.31. The dolomite formed from the solution

with added KCl has reaches a maximum cation ordering of 0.32. The addition of NaCl and KCl

both decrease the cation ordering of the dolomite after 800 hours in comparison to the stock

solution with no added salt.

26

CHAPTER V

DISCUSSION AND CONCLUSIONS

Three key observations help drive this research: (1) dolomite formed in peritidal settings

is more abundant and more stoichiometric than dolomite formed in subtidal settings (Lumsden

and Chimahusky, 1980), (2) dolomites formed in association with evaporite minerals are more

stoichiometric (Sass and Bein, 1988; Figure 6, 10, 11), and (3) numerous physicochemical

factors have been shown to impact dolomite stoichiometry (e.g., temperature, burial depth, water

depth, degree of evaporation, presence of [SO42-], fluid chemistry ([NaCl], [KCl], [Ca], and

[Mg]), and the Mg/Ca ratio in the fluid (Graf and Goldsmith, 1956; Glover and Sippel, 1967;

Land, 1985; Kaczmarek and Sibley, 2011; Kaczmarek and Thornton, 2017).

Comparison to Previous Work

Reaction Curve

The data presented in Figure 5 is consistent with previous data of Kaczmarek and Sibley

(2014). The general S-shaped pattern of the reaction curve and identification of four distinct

stages (induction, rapid replacement, primary recrystallization, and secondary recrystallization)

are observed in data from each long-term experiment conducted in this study. Induction time,

slope (rate), change in slope (increasing and decreasing rates), stoichiometry increases, and

cation ordering increases, are each observed over time and are consistent with previous work

(Kaczmarek and Sibley, 2014; Kaczmarek and Thornton, 2017). This characteristic reaction

curve provides additional information regarding trends that are consistent with previous work.

27

Rate of Dolomitization (Protodolomite)

The data presented here are remarkably consistent with previous ooid temperature data.

Figure 12 (modified from Kaczmarek and Thornton, 2017) is a linear cross plot of percent

protodolomite (no ordering reflections) as a function of reaction time. Kaczmarek and Thornton

(2017) followed a similar procedure in which they dolomitized aragonite ooids at various high

temperatures (160°C, 180°C, 200°C, 218°C, 235°C, and 250°C). The experiments in this study

were conducted at 215°C and the data points from the zero salt experiments at this temperature

plot directly between the 200°C and 218°C data from Kaczmarek and Thornton (2017) (black

dots). The consistency provides added confidence that the experiments were well controlled and

performed in a manner consistent with accepted methodologies.

Figure 12: Percent protodolomite (no ordering reflections) as a function of reaction time. Linear cross plot modified from Kaczmarek and Thornton (2017). The black data points represent experiments from this study. Figure on previous page (35).

y=0.87ln(x)- 0.65R²=0.97

y=0.65ln(x)- 0.53R²=0.97

y=0.52ln(x)- 0.49R²=0.97

y=0.50ln(x)- 0.80R²=0.98

y=0.38ln(x)- 0.94R²=0.96

y=0.14ln(x)- 0.49R²=0.91

y=0.50ln(x)- 0.64R²=0.97

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

1 10 100 1000

Percen

tProtodo

lomite

(noorde

ringrefle

ctions)

Time(Hours)

250

235

218

200

180

160

215

28

Ooid Size

Data from Martin et al. (2013) and Kaczmarek et al. (2015; 2016) that used ooids sieved

to a smaller size fraction (300-355 µm) were utilized in this comprehensive analysis. In

comparison to the ooids analyzed in this study (355-425 µm), the smaller ooid reactants exhibit a

slightly faster reaction rate than the larger ooids. This follows Zempolich and Baker (1993) as

dolomitization is described as a mimetic (fabric preserving) process in which, over time, the

dolomitization front will progressively move inwards as solution seeps into the inner layers of

the ooid. Inherently, smaller ooids will dolomitize at a faster reaction rate. Even though the ooids

used in this study were a larger fraction size, the reaction rate data is generally consistent for

both ooid sizes, which suggests that ooid size is not as strong of a control on reaction rate as is

temperature.

Calcite Dolomitization

The new short-term experimental data are also consistent with previously published data

from previous calcite reactant experiments (Martin et al., 2013; Kaczmarek et al., 2015, 2016),

which show that stoichiometry of the initial protodolomite products increases with salt

concentration in the fluid. The short-term experimental data presented here demonstrate an

increase in stoichiometry with each increasing NaCl and increasing KCl after only 5 hours.

In previous calcite experiments, cation ordering increased with increasing salt

concentration (Martin et al., 2013; Kaczmarek et al., 2015, 2016). In the current study, however,

there is little evidence that salinity has a large impact on cation ordering. The main differences

between the two studies are (i) the mineralogy of the reactant (calcite v. aragonite), and (ii) the

29

calcite study used a higher concentration of NaCl in the experimental fluids. There is no direct

evidence that the mineralogy of the reactant causes this difference.

In Martin et al. (2013) three sets of high-temperature experiments were conducted in

which calcite reacted with dolomitizing fluids with low, medium, and high salinity

concentrations. In these experiments, the low salinity fluid had 0 g of salt added, the medium

salinity fluid had 2.97g of NaCl added per 15 mL, and the high salinity fluid had 5.87 g of NaCl

added per 15 mL stock solution. In comparison, the long-term experiments conducted in this

study used zero salt, seawater, and hypersaline fluids with 0 g, 0.4251 g, and 2.57 g of NaCl

added per 15 mL stock solution, respectively. In comparing these sets of experiments, the low

salinity and medium salinity from Martin et al. (2013) correlate respectively with the zero salt

and hypersaline solutions from the experiments in this study. In Kaczmarek et al. (2013), very

little difference in cation ordering was observed between these two solutions. Similarly, very

little difference in cation ordering was observed in the NaCl solution within a range of 0-2.97 g

per 15 mL. The calcite data from Martin et al., (2013) suggests that at extremely high salt

concentrations (5.87 g/15 mL), cation ordering suddenly increases significantly to values higher

than is typically observed in laboratory dolomites (Kaczmarek and Sibley, 2014) and even higher

than natural dolomites (Manche and Kaczmarek, 2019a) and therefore may be an artifact of the

extremely high salt concentrations.

Mechanisms for Salt Influence

Fluid chemistry ([NaCl], [KCl], [Ca], and [Mg]) affects the kinetics of the dolomite

replacement reaction while the thermodynamic stability of the dolomite replacement reaction is a

function of the fluid Mg/Ca and temperature. Four proposed explanations for the change in

30

dolomitization rate and stoichiometry presented in this dataset include: (i) the competing ion

effect, (ii) solubility of the solid reactant, (iii) surface poisoning, and (iv) Mg-de-hydration.

Per the competing ion effect, as concentrations of competing ions in the fluid are

increased, the solution will lose energy and it becomes increasingly difficult to form highly

ordered minerals such as dolomite (Folk and Land, 1975). Instead, it becomes more likely that

minerals requiring less ordering will form due to the smaller amount of energy required. This can

explain the higher abundances of calcite and aragonite in marine-derived fluids, as these are

minerals with more simple structures than dolomite. In less concentrated solutions, there is less

interference of ions and slower rates of crystallization allows for highly ordered minerals such as

dolomite to form (Folk and Land, 1975). Folk and Land (1975) found that while hypersaline

solutions provide the necessary ions and a high Mg/Ca ratio, which both promote dolomite, it

must be significantly diluted to favor dolomite production and ordering.

Another way to explain the lower reaction rates and higher stoichiometries observed in

the salt experiments is how fluid chemistry can impact the solubility of the solid reactant. In the

scope of this project, as concentrations of NaCl, KCl, Mg, and Ca increase in the fluid, the

solubility of the aragonite ooid reactant will decrease (James and Jones, 2016). Specifically,

James and Jones (2016) discuss how an increase in magnesium ions in the fluid leads to a more

distorted crystal lattice. This also results in less carbonate ions being available in the solution and

decreases the rate of dolomitization due to a decreased availability of carbonate ions to form

dolomite. Dolomite is less soluble in solution because of its higher stability that comes from its

ordered structure and smaller magnesium ions in its crystal lattice in comparison to the calcium

ions in calcite. The findings presented here support the idea that the presence of certain cations

with strong hydration enthalpies appear to promote Mg dehydration in aqueous solutions, thus

31

aiding in the addition of Mg into protodolomite (e.g., Gaines, 1974; Vandeginste et al., 2019).

This explanation is supported by Debeye-Huckel theory, which shows that Mg-complexes are

less stable in solutions with higher ionic strength.

A poisoning effect may also explain how fluid chemistry in the experiments impacts

dolomitization rates and product composition. In this scenario, the surface of the aragonite ooid

reactants are poisoned by the adsorption of impurities and ionic compounds present in the

solution. Deelman (1999), for example, proposed that chlorine ions can easily bind with common

cations in marine fluids to produce compounds such as MgCl2, NaCl, and KCl that can sorb onto

mineral surfaces and thus decrease the reactive surface area of the calcium carbonate reactants.

This would decrease the number of surface sites available for the Mg to fit onto, which decreases

the rate of dolomitization (Deelman, 1999). This is a potential explanation of the observations

presented here as the rate of dolomitization does decrease with an increase in NaCl, KCl, MgCl2,

and CaCl2.

Lastly, Mg-de-hydration effects the dolomitization reaction by limiting the ability of Mg

to participate in the reaction (Lippmann, 1973). Because Ca ions have larger ionic radii than Mg

ions, they will more readily hydrate from the MgCl2 * 6H2O in the solution (Lippmann, 1973).

This results in the dehydrating of the Mg ions and causes either disordered calcium magnesium

carbonates or hydrous magnesium carbonates to form preferentially to dolomite (Lippmann,

1973). In the solutions with added Na and K, the Na and K ions also have larger ionic radii and

will even more readily hydrate from the MgCl2 * 6H2O in the solution. Additionally, as

temperatures increase, the dehydration of Mg will increase and a higher amount of Mg ions will

readily participate in the reaction (Lippman, 1973). Additionally, aqueous Mg ions are more

strongly hydrated than other divalent cations and will have a larger dehydration effect at higher

32

temperatures (Noyse, 1962; Lippman, 1973; Arvidson and Mackenzie, 1999; Stephenson et al.,

2008) As the Mg dehydrates, it will become more available to form into a carbonate lattice and

thus the rate of dolomitization will increase (Lippmann, 1973; Arvidson and Mackenzie, 1999).

Implications

The data presented here provide the basis for an alternative interpretation for the

relationship between evaporative settings and stoichiometry. Specifically, this dataset

demonstrates that precipitation of evaporites is not required for stoichiometric dolomite to form.

By altering the concentrations of major cations in dolomitizing fluids in this series of

experiments, more stoichiometric dolomite formed, without the need for invoking a high fluid

Mg/Ca associated with gypsum precipitation. These results show that fluid chemistry is an

important parameter effecting the rate and composition of the initial dolomite phase.

Dolomite as a Proxy for Understanding Fluid Chemistry

In previous work, evaporative conditions and recrystallization have been noted as the

primary cause of highly stoichiometric dolomites (Fuchtbauer and Goldschmidt, 1965; Schmidt,

1965; Sass and Katz, 1982; Langbein et al., 1984; Sperber et al., 1984; Kaczmarek and Sibley,

2011). These studies each observed that due to both thermodynamic and kinetic controls, higher

Mg/Ca ratios in the dolomitizing fluid resulted in higher amounts of magnesium in the product

dolomite. Glover and Sippel (1967), Kaczmarek at al. (2016), and Kaczmarek and Thornton

(2017) demonstrated with high-temperature experiments that in addition to a high Mg/Ca ratio in

the dolomitizing fluid, temperature can also directly impact the stoichiometry of evaporative

dolomites. Logan (1987) noted natural fluids with higher Mg/Ca ratios also tend to have

33

increased temperatures and salinity, which can exert additional pressure on the stoichiometry of

the product dolomite (Kaczmarek and Thornton, 2017; this study).

This study is part of an ongoing investigation to assess the controls on dolomite

stoichiometry (Kaczmarek and Sibley, 2011; Kaczmarek and Thornton, 2017; Manche and

Kaczmarek, 2019a). Manche and Kaczmarek (2019a) combined laboratory analyses with field

work by utilizing high resolution XRD data to assess stratigraphic variations in dolomite

stoichiometry in natural dolomites. In conjunction with Kaczmarek and Sibley (2011) and

Kaczmarek and Thornton (2017), this study provides a new way to explain the abundance of

stoichiometric dolomites in evaporitic depositional settings without the need for a high Mg/Ca

ratio or gypsum precipitation. This dataset provides evidence that salinity is an additional

parameter that needs to be considered as it directly affects the rate and composition of dolomite.

More stoichiometric dolomites, therefore, should not be solely attributed to increased Mg/Ca and

temperature alone. Instead, the concentrations of Na, K, Mg, and Ca in the dolomitizing fluid

must be considered as additional controlling factors.

Previous studies have proposed the use of stoichiometry as a useful proxy for conditions

of dolomitization (Lumsden and Chimahusky, 1980; Kaczmarek and Sibley, 2011; Ren and

Jones, 2017; Manche and Kaczmarek, 2019a). Lumsden and Chimahusky (1980) pointed out that

the compositional zonation commonly observed in ancient dolomite crystals suggests that

stoichiometry is controlled in part by the chemistry of the dolomitizing fluids. This coupled with

broad trends observed between dolomite stoichiometry and depth, Lumsden and Chimahusky

(1980) also posited that environmental factors and fluid flow may exert additional control on

stoichiometry of stratigraphically related dolomites. Kaczmarek and Sibley (2011) reasoned that

the as the Mg/Ca of the dolomitizing fluids decreased along a flow path (i.e. as dolomite forms,

34

Ca is released back into the fluids – Equation 1), that a decrease in dolomite abundance and

stoichiometry should result. For example, for hypersaline reflux, they suggested that one might

expect a decrease in stoichiometry down section as the fluids become progressively depleted in

Mg. Utilizing this concept, Ren and Jones (2017) cited observed trends in which dolomite

stoichiometry decreased laterally towards the center of Grand Cayman Island to interpret the

inward migration of seawater as the main control on both the rate of dolomitization and the

composition of the dolomite. They argued that as the seawater dolomitized the rock and

progressively lost Mg ions (i.e. decreased the Mg/Ca ratio in the fluids) as it moved toward the

island center, the dolomite produce was both less abundant and less stoichiometric. Most

recently, Manche and Kaczmarek (2019a) took this idea further and presented a high-resolution

vertical logs documenting patterns in dolomite abundance, dolomite stoichiometry, dolomite

crystal size, and stable oxygen isotopes for the Cretaceous dolomites of the Upper Glen Rose

Formation in central Texas. Although no variability was observed laterally on the scale of 100’s

of meters, transgressive and regressive vertical facies successions were observed to oscillate in

conjunction with dolomite abundance, stoichiometry, dolomite crystal size, and δ18O. These

parameters (except for dolomite crystal size) exhibited increases in transgressive facies

successions (Manche and Kaczmarek, 2019a). The opposite trends were observed in the

regressive facies successions. Manche and Kaczmarek (2019a) used these observations to

propose a model of syndepositional dolomitization in which patterns of dolomitization of

individual facies occur prior to the deposition of overlying sediments. Taken collectively, the

studies by Ren and Jones (2017) and Manche and Kaczmarek (2019a) demonstrate the

usefulness of stoichiometric variations in both the lateral and vertical directions to constrain the

conditions of dolomitization. Salt concentrations vary naturally in marine settings and are an

35

additional parameter that need to be considered. The experiments in this study address the

varying salt concentrations when interpreting patterns in stoichiometry. The new dataset

presented here is consistent with what has been observed in nature demonstrating the use of

stoichiometry as a proxy for conditions of dolomitization (Ren and Jones, 2017; Manche and

Kaczmarek, 2019a). These results also add to a growing body of research (Manche and

Kaczmarek, 2019a; Ryan et al., in review; Laya et al., in review) that suggests evaporite-

associated dolomites, though a common association, may not reflect a causal relationship

between highly evaporative (i.e. gypsum saturation) fluids and dolomitization.

Global Correlations

The findings presented here may have additional global implications. Certain

sedimentary basins, such as the Michigan Basin, which is a restrictive and periodically isolated

basin, were periodically evaporative and hence had higher salinity (Rine et al. 2017a; Rine et al.,

2017b). In a parallel study (Manche and Kaczmarek, 2019b), the global scale relationships

between dolomite stoichiometry and secular variations in seawater chemistry and depositional

mineralogy were examined in more detail. Manche and Kaczmarek (2019b) reported patterns of

dolomite abundance, stoichiometry, and δ18O that oscillated vertically with transgressive and

regressive facies successions. As an example, the data presented here may be useful as a

reference to interpret other times in Earth’s history. One example of this includes understanding

icehouse and greenhouse periods in Earth’s history. The global ocean exhibits trends of calcitic

seas during icehouse periods and aragonitic seas during greenhouse periods (Lowenstein et al.,

2001). Lowenstein et al. (2001) used fluid inclusions to interpret these oscillations in seawater

chemistry because of varied evaporite mineralogy. The aragonitic seas of the greenhouse periods

36

exhibit significantly lower salt concentrations in the seawater (Lowenstein et al., 2001). The

study of these varying seawater conditions can be combined with the additional consideration of

salt concentrations presented here to better understand the conditions of dolomitization in Earth’s

history.

Conclusions

The experimental data presented here are broadly consistent with the findings from

previous high-temperature dolomitization studies. For instance, these new data show that

dolomite replaces aragonite reactants in a stepwise process whereby aragonite is replaced by

protodolomite, which is replaced by poorly-ordered dolomite, which is replaced by well-ordered,

stoichiometric dolomite. The data presented here also show that higher concentrations of the

cations Na, K, Cl, Mg, and Ca, which are common constituents in marine fluids, significantly

increase dolomite stoichiometry. The concentrations of Mg and Ca also significantly increase

the rate of dolomitization, but the data for the Na and K experiments show only minor effects on

reaction rate. No effect on cation ordering was observed in the experiments. The finding that

cation concentrations exert positive pressure on dolomite stoichiometry and reaction rate

provides a new way to explain the common observations of higher dolomite abundances and

more stoichiometric dolomites interpreted to have formed in evaporitic settings. Previous studies

have invoked gypsum precipitation during evaporative conditions to cause higher fluid Mg/Ca

ratios, which in turn increase reaction rates and exert positive pressure on dolomite

stoichiometry. These new data show that the conventional explanation may be more nuanced that

previously accepted. Specifically, that higher Mg/Ca ratios driven by evaporite precipitation may

not be needed for higher stoichiometry; only elevated concentrations of common cations. In

37

addition to temperature, Mg/Ca, reactant mineralogy, organic matter, and sulfate, the

concentrations of Na, K, Cl, Mg, and Ca should be considered as additional variables when

interpreting temporal and spatial patterns in dolomite stoichiometry. The observations presented

here add to our understanding of the fundamental controls on dolomite stoichiometry and

reaction rate, and can help further constrain geological interpretations based on dolomite

stoichiometry, which has recently been shown to be a useful proxy in our effort to elucidate the

origin of ancient dolomites.

38

REFERENCES

Arvidson, R.S., Mackenzie, F.T. (1999) The dolomite problem: control of precipitation kinetics by temperature and saturation state. Am. J. Sci. 299, 257–288.

Baker, P.A. and Kastner, M. (1981) Constraints on the Formation of Sedimentary Dolomite: Science, v.213, p.214–216.

Budd, D.A. (1997) Cenozoic dolomites of carbonate islands: their attributes and origin. Earth

Sci. Rev. 42, 1–47.

Carpenter, A. (1980) The chemistry of dolomite formation I: the stability of dolomite. In: Zenger, D.H., Dunham, J.B., Ethington, R.L. (Eds.), Concepts and Models of Dolomitization. SEPM Spec. Publ. 28. pp. 111–121.

Cohen, H.F. and Kaczmarek, S.E. (2017) Evaluating the effects of fluid NaCl and KCl concentrations on reaction rate, major cation composition, and cation ordering during high-temperature dolomitization experiments (poster), Geological Society of America 2017 Annual Meeting, Oct. 22-25, 2017, Seattle, WA, #306364

Cohen, H. and Kaczmarek, S.E. (2018) Evaluating the effects of fluid chemistry on dolomite stoichiometry and reaction rate (poster), American Association of Petroleum Geologists Annual Conference, May 20-23, Salt lake City, UT, #2855811.

Cole-Parmer Instrument Company, LLC. (2019) Cole-Parmer Mortar and Pestle Sets, Agate, https://www.coleparmer.com/p/cole-parmer-mortar-and-pestle-sets-agate/64980.

de Dolomieu, D.G. (1791) Sur un genre de pierres calcaires trés peu effervescentes avec les acides et phosphorescentes par la collision (on a types of calcareous rock that reacts very slightly with acid and that phosphoresces on being struck). J. Physique 39, 3–10.

Deelman, J. C. (1999) Low-temperature nucleation of magnesite and dolomite - Neues Jahrbuch für Mineralogie, Monatshefte, 1999 (7): 289-302, Stuttgart 1999.

DWK Life Sciences LLC. (2017) Desiccators: Borosilicate Desiccator with Stopcock. https://www.kimble-chase.com/advancedwebpage.aspx?cg=3483&cd=5&WebID=635&SK U=644&SKUFLD=SKU&SKUTYPE=

Folk, R.L. and Land, L.S. (1975) Mg/Ca Ratio and Salinity: Two Controls over Crystallization of Dolomite: AAPG Bulletin, v.59, no.1, p.60–68.

Fuchtbauer, H., Goldschmidt, H. (1965) Beziehungen zwischen calciumgehalt und bil- dungs-

bedinggungen der dolomite. Geol. Rundsch. 55, 29–40.

Gabellone, T. and Whitaker, F. (2016) Secular variations in seawater chemistry controlling dolomitization in shallow reflux systems: insights from reactive transport modelling: Sedimentology, 27 p., doi: 10.1111/sed.12259.

39

Gaines, A.M. (1974) Protodolomite synthesis at 100 C and atmospheric pressure, Science, 183, p. 518-520

Glover, E.D., and Sippel, R.F. (1967) Synthesis of magnesium calcites: Geochimica et

Cosmochimica Acta, v. 31, i. 4, pp. 603–613. Goldsmith, J.R., and Graf, D.L. (1958) Structural and compositional variations in some natural

dolomites: J. Geol., v. 66, p. 678–693. Graf, D.L., and Goldsmith, J.R. (1956) Some hydrothermal synthesis of dolomite and proto-

dolomite: J. Geol., v. 64, p. 173–187. Gregg, J.M., Bish, D.L., Kaczmarek, S.E., and Machel, H.G. (2015) Mineralogy, nucleation, and

growth of dolomite in the laboratory and sedimentary environment: a review: Sedimentology, v.62, p.1749–1769.

Illing, L.V., Wells, A.J., Taylor, J.C.M. (1965) Penecontemporary dolomite in the Persian Gulf.

In: Pray, L.C., Murray, R.C. (Eds.), Dolomitization and Limestone Diagenesis: Soc. Econ. Paleontologists and Mineralogists Spec. Pub Vol. 13. pp. 89–111.

James, N.P. and Jones, B. (2016) Origin of Carbonate Sedimentary Rocks: Wiley, 464 p. Jones, B., Luth, R.W., MacNeil, A.J. (2001) Powder X-ray analysis of homogeneous and

heterogeneous dolostones. J. Sediment. Res. 71, 791–800.

Jones, B., Luth, R.W. (2002) Dolostones from Grand Cayman, British West Indies. J. Sediment. Res. 72, 560–570.

Jones, G.D. and Xiao, Y. (2005) Dolomitization, anhydrite cementation, and porosity evolution in a reflux system: Insidghts from reactive transport models: AAPG Bulletin, v.89, no.5, p.577–601, doi: 10.1306/12010404078.

Kaczmarek, S.E. (2005) Crystal growth mechanisms in natural and synthetic dolomite: Insight

into dolomitization kinetics, Ph.D. Dissertation, Michigan State University, 230 p. Kaczmarek, S.E., and Sibley, D.F. (2007) A Comparison of Nanometer-Scale Growth and

Dissolution Features on Natural and Synthetic Dolomite Crystals: Implications for the Origin of Dolomite: Journal of Sedimentary Research, v. 77, p. 424–432, doi: 10.2110/jsr.2007.035.

Kaczmarek, S.E., and Sibley, D.F. (2011) On the evolution of dolomite stoichiometry and cation

order during high-temperature synthesis experiments: An alternative model for the geochemical evolution of natural dolomites: Sedimentary Geology, v. 240, p. 30–40, doi: 10.1016/j.sedgeo.2011.07.003.

Kaczmarek, S.E., and Sibley, D.F. (2014) Direct physical evidence of dolomite recrystallization:

Sedimentology, v. 61, p. 1862–1882, doi: 10.1111/sed.12119.

40

Kaczmarek, S.E. and Thornton, B. (2015) Development and evolution of cation order during dolomitization (poster), 15th Bathurst Meeting of Carbonate Sedimentologists, July 13-16, University of Edinburgh, U.K.

Kaczmarek, S.E., Whitaker, F.F., Avila, M., Lewis, D., and Saccocia, P.J. (2016) A case for

caution when using geochemical models to make predictions about dolomite (oral), AAPG/SEPM Annual Meeting, June 20-22, Calgary, Canada.

Kaczmarek, S.E., and Thornton, B. (2017) The effect of temperature on stoichiometry, cation

ordering, and reaction rate in high-temperature dolomitization experiments: Chemical Geology.

Kaczmarek, S.E., Whitaker, F.F., Avila, M., Lewis, D., and Saccocia, P.J. (2015) A case for

caution when using geochemical models to make predictions about dolomite (oral), AAPG/SEPM Annual Meeting, June 20–22, Calgary, Canada.

Kaczmarek, S.E., Gregg, J.M., Bish, D.L., Machel, H.G., and Fouke, B.W. (2017) Dolomite,

very high-magnesium calcite, and microbes—Implications for the microbial model of dolomitization, in MacNeil, A., Lonnee, J., and Wood, R., eds., Characterization and Modeling of Carbonates— Mountjoy Symposium 1: Society for Sedimentary Geology (SEPM) Special Publication 109, p. 7–20, http://dx.doi.org/10.2110/sepmsp.109.01.

Land, L.S. (1967) Diagenesis of skeletal carbonates. J. Sediment. Petrol. 37, 914–930.

Land, L. (1980) The isotopic and trace element geo- chemistry of dolomite: The state of the art, in Zenger, D.H., Dunham, J.B. and Ethington, R.L., eds., Concepts and Models of Dolomitization: Society of Economic Paleontologists and Mineralogists (SEPM) Special Publication 28, p. 87–110, https://doi.org/10.2110/pec.80.28.0087.

Land, L. (1985) The Origin of Massive Dolomite: Journal of Geoscience Education, v.33, p.112–125, https://doi.org/10.5408/0022-1368-33.2.112.

Land, L.S. (1998) Failure to precipitate dolomite at 25 °C from dilute solution despite 1000-fold

oversaturation after 32 years. Aquat. Geochem. 4, 361–368.

Langbein, R. Von, Landgraf, K.F., Milbrodt, E. (1984) Calcium uberschusse im dolomit als indikator des sedimentationsmilieus in devonischen karbonatgesteinen. Chem. Erde 43, 217–227.

Laya, J.C., Teoh, C.P., Whitaker, F., Manche, C., Kaczmarek, S.E., Tucker, M., Gabellone, T. (in review) Dolomitization of a Miocene-Pliocene progradational carbonate platform: revising reflux model on Bonaire Island, Journal of Sedimentary Research.

Liebermann, O. (1967) Synthesis of Dolomite: Nature, p. 241–245. Lippmann, F. (1973) Sedimentary carbonate minerals: New York, Springer-Verlag, 228 p.

41

Logan, B.W. (1987) The MacLeod evaporite basin, Western Australia: AAPG Memoir 44, 140 p.

Lowenstein, T.K., Timofeef, M.N., Brennan, S.T., Hardie, L.A., and Demicco, R.V. (2001)

Oscillations in Phanerozoic Seawater Chemistry: Evidence from Fluid Inclusions: Science, v. 294, pp. 1086–1088.

Lumsden, D.N. (1979) Discrepancy between thin-section and X-ray estimates of dolomite in

limestone: Journal of Sedimentary Research, v. 49, p. 429–436, doi:10.1306/212F7761-2B24-11D7-8648000102C1865D.

Lumsden, D.N., Chimahusky, J.S. (1980) Relationship between dolomite non-stoichiometry and

carbonate facies parameters. In: Zenger, D.H., Dunham, J.B., Ethington, R.L. (Eds.), Concepts and Models of Dolomitization. SEPM Spec. Publ. 28. pp. 123–137.

Machel, H.G. (2004) Concepts and models of dolomitization: a critical reappraisal. In: Braithwaite, C. J. R., Rizzi G., and Drake, G. (eds) The Geometry and Petrogenesis of Dolomite Hydrocarbon Reservoirs: Geological Society, London, Special Publications, v. 235, p. 7–63, doi: 10.1144/GSL.SP.2004.235.01.02.

Malone, M.J., Baker, P.A., Burns, S.J. (1996) Recrystallization of dolomite: an experimental

study from 50–200 °C. Geochim. Cosmochim. Ac. 60, 2189–2207.

Manche, C.J. and Kaczmarek, S.E. (2019a) Evaluating reflux dolomitization using a novel high-resolution record of dolomite stoichiometry: A case study from the Cretaceous of Central Texas, U.S.A., Geology, v. 47, p. 586-590, doi.org/10.1130/G46218.1.

Manche, C. and Kaczmarek, S.E. (2019b) Evaluating the relationship between stoichiometry and

cation ordering in ancient dolomites (poster) American Association of Petroleum Geologists Annual Conference, May 19-22, San Antonio, TX

Martin, S.L., Cudnik, B., DeMille, J., Montoya, C., Barrie, E. and Kaczmarek, S.E. (2013) The

effect of salinity on the rate of dolomitization (poster), 2013 AAPG Annual Meeting, Pittsburg, PA.

Morrow, D.W. and Ricketts, B.D. (1988) Experimental investigation of sulfate inhibition of

dolomite and its mineral analogues: Shukla, Vijai, Baker, Paul A. Special Publication Society of Economic Paleontologists and Mineralogists, v. 43, p. 25–38

Noyse, R.M. (1962) Thermodynamics of ion hydration as a measure of effective dielectric

properties of water. J. Am. Chem. Soc. 84, 513–522.

Parr Instrument Company (2019) Sample Preparation Bombs: Bulletin 4700, http://www.instrumentalia.com.ar/uploads/archivos/4700MB_Parr.pdf.

42

Reeder, R.J. (1983) Crystal chemistry of the rhombohedral carbonates. In: Reeder, R.J. (Ed.), Carbonates: mineralogy and chemistry. Mineralogical Society of America Reviews in Mineralogy 11. pp. 1–47.

Reeder, R.J. (2000) Constraints on cation order in calcium-rich sedimentary dolomite. Aquat. Geochem. 6, 213–226.

Reeder, R.J., and Sheppard, C.E. (1984) Variation of lattice parameters in some sedimentary dolomites: American Mineralogist, v. 69, no. 5–6, p. 520–527.

Ren, M. and Jones, B. (2017) Spatial variations in the stoichiometry and geochemistry of

Miocene dolomite from Grand Cayman: implications for the origin of island dolostone: Sed. Geol. 348, p. 69–93.

Rine, M., Garrett, J., and Kaczmarek, S.E. (2017a) A new facies architecture model for the

Silurian Niagara-Lower Salina “Pinnacle” Reef Complexes of the Michigan Basin, in: Characterization and Modeling of Carbonates - Mountjoy Symposium 1 (Eds. A. MacNeil, J. Lonnee, & R. Wood) SEPM Special Publication 109, p. 70-86, doi: doi.org/10.2110/sepmsp.109.02.

Rine, M.J., Kaczmarek, S.E., and Garrett, J.D. (2017b) A new sequence stratigraphic model for

the Silurian A-1 Carbonate (Ruff Formation) of the Michigan Basin, in: Paleozoic Stratigraphy and Resources of the Michigan Basin: Geological Society of America Special Paper 531 (Eds. G.M. Grammer, W.B. Harrison, III, and D.A. Barnes), p. 175-195, doi:10.1130/2017.2531(09).

Rose, K. and Kaczmarek, S.E. (2019) Effects of reactant surface area and mineralogy on

dolomite stoichiometry, AGU Annual Conference, San Francisco, Dec. 2019 Royse, C.F., Wadell, J.S., and Petersen, L.E. (1971) X-ray determination of calcite-dolomite; an

evaluation: Journal of Sedimentary Research, v. 41, p. 483–488, doi: 10.1306/74D722A7-2B21-11D7-8648000102C1865D.

Ryan, B.H., Kaczmarek, S.E., and Rivers, J. (2019) Dolomite dissolution: An alternative

pathway for the formation of palygorskite clay, Sedimentology, v. 66, p. 1803-1824, doi.org/10.1111/sed.12559.

Ryan, B., Kaczmarek, S.E., and Rivers, J. (in review) A reassessment of evaporite-associated

peritidal dolomites: How robust is the hypersaline reflux dolomitization model? Sedimentology

Sass, E., and Bein, A. (1988) Dolomites and salinity: A comparative geochemical study, in

Shukla, V., and Baker, P., eds., Sedimentology and Geochemistry of Dolostones: Society of Economic Paleontologists and Mineralogists Special Publication, v. 43, p. 223–233, https://doi.org /10.2110/pec.88.43.0223.

Sass, E., Katz, A. (1982) The origin of platform dolomites. Am. Jour. Sci. 282, 1184–1213.

43

Schmidt, V. (1965) Facies, diagenesis and related reservoir properties in the Gigas beds (upper Jurassic), northwestern Germany. In: Pray, L.C., Murray, R.C. (Eds.), Dolomitization and Limestone Diagenesis. Soc. Econ. Paleont. Mineral. Spec. Pub. 13. pp. 124–168.

Searl, A. (1994) Discontinuous solid solution in Ca-rich dolomites: the evidence and implications for the interpretation of dolomitic petrographic and geochemical data. In: Purser, B., Tucker, M., Zenger, D. (Eds.), Dolomites: A Volume in Honour of Dolomieu. International Association of Sedimentologists, Special Publication, pp. 361–376.

Shinn, E., Ginsburg, R.N., Lloyd, R.M. (1965) Recent supratidal dolomite from Andros Island. In: Pray, C., Murray, R.C. (Eds.), Dolomitization and Limestone Diagenesis. SEPM Spec. Publ. 13. pp. 112–123.

Sibley, D.F. (1990) Unstable to stable transformations during dolomitization: The Journal of Geology, v. 98, p. 739–748, https://doi.org/10.1086 /629437.

Sibley, D.F., Dedoes, R.E., and Bartlett, T.R. (1987) Kinetics of dolomitization: Geology, v.15, p.1112–1114.

Sibley, D.F., Nordeng, S.H., Borkowski, M.L. (1994) Dolomitization kinetics in hydrothermal

bombs and natural settings. J. Sediment. Res. A64, 630–637.

Sperber, C.M., Wilkinson, B.H., Peacor, D.R. (1984) Rock composition, dolomite stoichiometry, and rock water reactions in dolomitic carbonate rocks. J. Geol. 92 (6), 609–622.

Stephenson, A.E., DeYoreo, J.J., Wu, L., Wu, K.J., Hoyer, J., Dove, P.M. (2008) Peptides enhance magnesium signature in calcite: insights into origins of vital effects. Science 322, 724–727.

Sun, S.Q. (1994) A Reappraisal of Dolomite Abundance and Occurrence in the Phanerozoic: Journal of Sedimentary Research, v. A64, n. 2, p. 396–404.

Thermo Fisher Scientific Inc. (2012) Thermo Scientific Heratherm General Protocol: Operating

Instructions, https://assets.fishersci.com/TFS-Assets/LED/manuals/D21463~.pdf.

Thornton, B. and Kaczmarek, S.E. (2015) Stepwise reaction pathway and temperature-dependence of dolomitization of aragonite ooids (poster), 2015 AAPG/SEPM Annual Meeting, May 31 – June 3, Denver, CO.

Turpin, M., Nader, F.H., Kohler, E. (2012) Empirical calibration for dolomite stoichiometry

calculation: application on Triassic muschelkalk-letternkohle carbonates (French Jura). Oil and Gas Science and Technology – Reviews IFP Energy nouvelles 67, 77–95.

Usdowski, E. (1994) Synthesis of dolomite and geochemical implications. In Dolomites: A Volume in Honour of Dolomieu, B.H. Purser, v. 21, p. 345–360, doi: 10.1002/9781444304077.

Van Tuyl, F.M. (1916) New points on the origin of dolomite. Am. J. Sci. 42, 249–260.

44

Vandeginste, V., Snell, O., Hall, .M.R., Steer, E., and Vandeginste, A. (2019) Acceleration of dolomitization by zinc in saline waters, Nature Communications, 10, 8 p., article number: 1851.

Warren, J. (2000) Dolomite: occurrence, evolution and economically important associations: Earth-Science Reviews, v. 52, p. 1–81.

Zempolich, W.G., and Baker, P.A. (1993) Experimental and Natural Mimetic Dolomitization of

Aragonite Ooids: Journal of Sedimentary Research, v. Vol. 63, p. 596–606, doi: 10.1306/D4267B86-2B26-11D7-8648000102C1865D.

Zhang, F., Xu, Xu, H., Konishi, H., and Roden, E.E. (2010) A relationship between d104 value

and composition in the calcite-disordered dolomite solid-solution series: American Mineralogist, v. 95, no. 11–12, p. 1650–1656.

Zhang, F., Xu, H., Konishi, H., Shelobolina, E.S., and Roden, E.E. (2012) Polysaccharide-

catalyzed nucleation and growth of disordered dolomite: A potential precursor of sedimentary dolomite: The American Mineralogist, v. 97, no. 4, p. 556–567, https://doi.org/10.2138/am.2012.3979.

Zhao, H., Jones, B. (2012) Origin of “island dolostones”: a case study from the Cayman formation (Miocene), Cayman Brac, British West Indies. Sediment. Geol. 243-244, 191–206.

45

Appendix A

Dataset

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

69

70

71