Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate...

40
Calcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu ¨ ttge Department of Oceanography, MS-3146, Texas A&M University, College Station, Texas 77843, and Department of Earth Science MS-126, Rice University, 6100 Main Street, P.O. Box 1892, Houston, Texas 77251 Contents 1. Introduction 342 2. Major Marine Carbonate Minerals 344 2.1. General Considerations 344 2.2. Physical and Chemical Properties 344 2.2.1. General Considerations 344 2.2.2. R3 hc Carbonates: Calcite, Magnesian Calcite, and Related Phases 345 2.2.3. R3 h Carbonates: Dolomite and Other Double Carbonates 347 2.2.4. Orthorhombic Carbonates: Aragonite 348 3. Carbonate Mineral Precipitation and Dissolution Kinetics Relevant to Seawater 348 3.1. General Overview 348 3.2. Mineral Reactivity and the Mineral-Solution Interface: The Link Between Thermodynamic Driving Force and the Overall Rate 349 3.2.1. General Rate Equations 349 3.2.2. Relationship to Mineral Surface Processes 350 3.2.3. Treatment of Kinetic Inhibition 351 3.2.4. Mineral Reaction Rates and Surface Area 352 3.2.5. Dependence on Distance from Equilibrium and Scaling 353 3.3. Precipitation of Carbonate Minerals from Seawater and Related Solutions 353 3.3.1. General Remarks 353 3.3.2. Saturation State Controls and Their Relationship to Inhibition 354 3.3.3. The Effect of Salinity and Ionic Strength 354 3.3.4. The Role of Temperature 355 3.3.5. The Roles of Magnesium and Sulfate 355 3.3.6. The Role of Other Impurities: Manganese, Strontium, Phosphate, and Organic Matter 356 3.3.7. Nucleation of Calcium Carbonate 357 3.4. Dissolution Reactions 359 3.4.1. Calcite Dissolution Kinetics 359 3.4.2. Aragonite Dissolution 360 3.4.3. Magnesian Calcite Dissolution Kinetics 360 3.5. Biominerals and Nanophase Carbonates 361 3.5.1. General Considerations 361 3.5.2. Biomineralization 361 3.5.3. Biogeochemistry, Geomicrobiology 361 3.5.4. Nanophase Carbonates 362 3.6. Calcium Carbonate-Organic Matter Interactions 363 3.6.1. Overview 363 3.6.2. Indirect Evidence of the Influences of Organic Matter on Calcium Carbonate 363 3.6.3. Direct Evidence for Organic Matter Adsorption on Calcium Carbonate 365 4. Oceanic Sources of Marine Carbonates 366 4.1. Overview 366 4.2. Primary Carbonate Sources for Deep Sea Sediments 366 4.3. Shallow Water Biogenic Carbonate Sources 367 4.3.1. General Considerations 367 4.3.2. Sources of Fine-Grained Carbonate Sediments 367 4.3.3. Sources of Carbonate Sands 368 4.3.4. Carbonate Cements 370 4.3.5. Dolomite Formation 371 4.3.6. Carbonate Minerals Formed in Siliciclastic Sediments 371 5. Dissolution of Calcium Carbonate 372 5.1. Dissolution of Carbonate Minerals in the Pelagic Environment 372 5.1.1. General Considerations 372 5.1.2. Dissolution in the Water Column 372 5.1.3. Dissolution on the Deep Sea Floor 372 5.2. Dissolution of Carbonate Minerals in Carbonate-Rich Shallow Water Sediments 373 6. Response of Carbonate-Rich Sediments to the Acidification of the Ocean Due to Rising Atmospheric pCO 2 374 6.1. General Principles and Considerations 374 6.2. The Major Ocean Basins 374 6.3. Shallow Water Carbonate-Rich Sediments 375 7. Summary and Thoughts on Possible Future Research Directions 375 7.1. Summary 375 7.2. Thoughts on Possible Future Research Directions 377 8. Acknowledgements 378 9. References 378 1. Introduction Calcium carbonate is an important and often dominant component of marine sediments. This is reflected by the observation that about 20% by volume of Phanerozoic (0- 547 Ma; Ma ) million years ago) sedimentary rock is made up of the carbonate minerals calcite (CaCO 3 ) and dolomite (CaMg(CO 3 ) 2 ). 1 Over the history of the Earth, the primary source of marine carbonate minerals has shifted from abiotic precipitation to biogenic sources. The biogenic sources have in turn shifted from being primarily relatively shallow water benthic organisms to the current situation where small open ocean pelagic organisms, living mainly in the photic zone, dominate calcium carbonate formation. Much of the calcium * To whom correspondence should be addressed. Phone: (979)-845-9630. Fax: (979) 845-9631. E-mail: [email protected]. 342 Chem. Rev. 2007, 107, 342-381 10.1021/cr050358j CCC: $65.00 © 2007 American Chemical Society Published on Web 01/30/2007

Transcript of Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate...

Page 1: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

Calcium Carbonate Formation and DissolutionJohn W. Morse,* Rolf S. Arvidson, and Andreas Luttge

Department of Oceanography, MS-3146, Texas A&M University, College Station, Texas 77843, and Department of Earth Science MS-126,Rice University, 6100 Main Street, P.O. Box 1892, Houston, Texas 77251

Contents1. Introduction 3422. Major Marine Carbonate Minerals 344

2.1. General Considerations 3442.2. Physical and Chemical Properties 344

2.2.1. General Considerations 3442.2.2. R3hc Carbonates: Calcite, Magnesian

Calcite, and Related Phases345

2.2.3. R3h Carbonates: Dolomite and OtherDouble Carbonates

347

2.2.4. Orthorhombic Carbonates: Aragonite 3483. Carbonate Mineral Precipitation and Dissolution

Kinetics Relevant to Seawater348

3.1. General Overview 3483.2. Mineral Reactivity and the Mineral−Solution

Interface: The Link Between ThermodynamicDriving Force and the Overall Rate

349

3.2.1. General Rate Equations 3493.2.2. Relationship to Mineral Surface

Processes350

3.2.3. Treatment of Kinetic Inhibition 3513.2.4. Mineral Reaction Rates and Surface Area 3523.2.5. Dependence on Distance from Equilibrium

and Scaling353

3.3. Precipitation of Carbonate Minerals fromSeawater and Related Solutions

353

3.3.1. General Remarks 3533.3.2. Saturation State Controls and Their

Relationship to Inhibition354

3.3.3. The Effect of Salinity and Ionic Strength 3543.3.4. The Role of Temperature 3553.3.5. The Roles of Magnesium and Sulfate 3553.3.6. The Role of Other Impurities:

Manganese, Strontium, Phosphate, andOrganic Matter

356

3.3.7. Nucleation of Calcium Carbonate 3573.4. Dissolution Reactions 359

3.4.1. Calcite Dissolution Kinetics 3593.4.2. Aragonite Dissolution 3603.4.3. Magnesian Calcite Dissolution Kinetics 360

3.5. Biominerals and Nanophase Carbonates 3613.5.1. General Considerations 3613.5.2. Biomineralization 3613.5.3. Biogeochemistry, Geomicrobiology 3613.5.4. Nanophase Carbonates 362

3.6. Calcium Carbonate−Organic MatterInteractions

363

3.6.1. Overview 363

3.6.2. Indirect Evidence of the Influences ofOrganic Matter on Calcium Carbonate

363

3.6.3. Direct Evidence for Organic MatterAdsorption on Calcium Carbonate

365

4. Oceanic Sources of Marine Carbonates 3664.1. Overview 3664.2. Primary Carbonate Sources for Deep Sea

Sediments366

4.3. Shallow Water Biogenic Carbonate Sources 3674.3.1. General Considerations 3674.3.2. Sources of Fine-Grained Carbonate

Sediments367

4.3.3. Sources of Carbonate Sands 3684.3.4. Carbonate Cements 3704.3.5. Dolomite Formation 3714.3.6. Carbonate Minerals Formed in Siliciclastic

Sediments371

5. Dissolution of Calcium Carbonate 3725.1. Dissolution of Carbonate Minerals in the

Pelagic Environment372

5.1.1. General Considerations 3725.1.2. Dissolution in the Water Column 3725.1.3. Dissolution on the Deep Sea Floor 372

5.2. Dissolution of Carbonate Minerals inCarbonate-Rich Shallow Water Sediments

373

6. Response of Carbonate-Rich Sediments to theAcidification of the Ocean Due to RisingAtmospheric pCO2

374

6.1. General Principles and Considerations 3746.2. The Major Ocean Basins 3746.3. Shallow Water Carbonate-Rich Sediments 375

7. Summary and Thoughts on Possible FutureResearch Directions

375

7.1. Summary 3757.2. Thoughts on Possible Future Research

Directions377

8. Acknowledgements 3789. References 378

1. IntroductionCalcium carbonate is an important and often dominant

component of marine sediments. This is reflected by theobservation that about 20% by volume of Phanerozoic (0-547 Ma; Ma) million years ago) sedimentary rock is madeup of the carbonate minerals calcite (CaCO3) and dolomite(CaMg(CO3)2).1 Over the history of the Earth, the primarysource of marine carbonate minerals has shifted from abioticprecipitation to biogenic sources. The biogenic sources havein turn shifted from being primarily relatively shallow waterbenthic organisms to the current situation where small openocean pelagic organisms, living mainly in the photic zone,dominate calcium carbonate formation. Much of the calcium

* To whom correspondence should be addressed. Phone: (979)-845-9630.Fax: (979) 845-9631. E-mail: [email protected].

342 Chem. Rev. 2007, 107, 342−381

10.1021/cr050358j CCC: $65.00 © 2007 American Chemical SocietyPublished on Web 01/30/2007

Page 2: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

carbonate that is formed is dissolved in the water columnand, more importantly, in sediments through processescollectively known as diagenesis.

The occurrence of carbonate minerals in modern marinesediments can readily be divided into those found in shoal-to-shallow (0 to less than∼200 m) and deep water (generally>1 km) environments associated with the world’s majorocean basins. The factors controlling the sources, mineralogy,and diagenesis of carbonates in these environments are verydifferent. Within the shoal-to-shallow water environment, thesources and diagenesis of carbonates differ substantiallybetween carbonate-rich and primarily siliciclastic (composedmainly of detrital minerals such as clays and quartz sands)sediments.

Almost all deep sea carbonate-rich sediments are com-posed of calcite low in magnesium (>99% CaCO3). Thismaterial is primarily derived from pelagic skeletal organisms,with coccolithophores (plants) usually being the mostimportant quantitatively, followed by foraminifera (animals).In sediments overlain by waters of intermediate depth suchas mid-ocean ridge crests, aragonite derived from pelagicpteropods and heteropods can be found. Calcite cementscontaining abundant magnesium can also occur in deep watersediments, but they are relatively rare. Since the time of theHMS Challenger expeditions in the late 19th century, it hasbeen recognized that that the saturation state of seawateroverlying sediments in the deep ocean exerts a majorinfluence on the distribution and abundance of calciumcarbonate in these sediments

Shoal-to-shallow water, carbonate-rich sediments arelargely confined today to the subtropic and tropic climaticzones but are found even at high latitudes. These carbonatesare primarily produced by the disintegration of the skeletonsof benthic organisms, such as corals, echinoids, mollusks,benthic foraminifera, and coralline algae. In some environ-ments, inorganic precipitates such as cements and oo¨ids arealso abundant. The source of aragonite needle muds remainscontroversial. In siliciclastic sediments, the major source ofcarbonates is also primarily from benthic organisms. Theseinclude bivalves, sea urchins, and foraminifera.

Interest in calcium carbonate formation and dissolutionin the ocean has increased because of the central role thesereactions will play in the ocean’s response to the increasingpartial pressure of carbon dioxide (pCO2) in the atmosphere.This is causing acidification of the ocean. Reactions resultingin the formation and dissolution of calcium carbonate arealso of central importance to the sequestration of carbondioxide in subsurface carbonate reservoirs and saline waters.

The objective of this paper is to review the major aspectsof the chemistry of carbonate mineral formation and dis-

John Morse is originally from Minnesota where he received his B.S. fromthe University of Minnesota’s Institute of Technology in Geology (1969).From there he went to graduate school at Yale University where, underthe mentorship of Robert Berner, he received both his M.S. (1971) andPh.D. (1973) degrees in Geology. In 1973, he joined the faculty at FloridaState University where he held a joint appointment in the Oceanographyand Chemistry departments. Subsequently, he served on the faculty ofthe University of Miami (Florida) from 1976−1981 and then went to TexasA&M University where he currently holds the Louis and Elizabeth ScherckChair in Oceanography. His primary research interest has been in thephysical chemistry of sedimentary minerals and their nonequilibriuminteractions with natural waters.

Rolf Arvidson received a B.S. in Geology from the University of Washington(1980). After working as an exploration geologist in Denver, he receiveda M.S. in Geology from the University of Iowa under Philip H. Heckel anda Ph.D. in Oceanography with Fred T. Mackenzie in 1997. He collaboratedwith John Morse as a Postdoctoral Researcher at Texas A&M University(1997−1999), and at Rice University with Andreas Luttge (1999−2001).He is currently Research Scientist at Rice. His research interests includeexperimental, theoretical, and modeling approaches to mineral precipitationand dissolution kinetics, as well as integration of this understanding toboth short- and long-term biogeochemical cycling.

Andreas Luttge started his scientific education in Braunschweig (Germany)in 1979. In 1982, he went to Tubingen (Germany) where he got his Ph.D.in 1990 and his “Habilitation” in 1995 (Eberhard-Karls Universitat). Next,he followed an invitation from Yale University where he worked first as aHumboldt Fellow and later as an associate research scientist. In 1998,Luttge accepted an offer from Rice University in Houston. Currently, heholds a double appointment in Earth Science and Chemistry at Rice. Hisresearch interest is focused on the processes that govern fluid/solidinteractions from low-temperature conditions to the pressures andtemperatures of the deep crust. He is particularly interested in theparticipation of microorganisms in these processes. His work includesexperimental and modeling techniques that he applies to questions ofmineral reactions in sedimentary basins, weathering, the fate of nano-particles in the environment, atmospheric and global change, and microbialfuel cells.

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 343

Page 3: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

solution in the marine environment. Because the primaryreadership of this journal is among the chemical, rather thanEarth science community, we will put our primary effort intodealing with the chemistry of important marine carbonateminerals and their interaction with seawater. In doing so,we will not attempt to cover all that is known about carbonatemineral-aqueous solution chemistry, but rather focus pri-marily on those aspects of direct relevance to their behaviorin seawater.

Following this chemical review (sections 2 and 3), section4 will deal with observations of processes controlling thesources of carbonate minerals in the ocean, and section 5will address the processes leading to their dissolution; thefate of a major fraction of the carbonate minerals produced.Section 6 will focus on the likely responses of oceaniccarbonate minerals, in different marine environments, to theacidification of the oceans because of the central importanceof this response to impact of increased pCO2 on the marineenvironment and climate change. It should be noted that areview paper on the carbonic acid system in seawater byMillero2 also appears in this issue. We will therefore notcover this topic in any detail in this paper. However, theMillero paper should be referred to because the behavior ofthe carbonic acid system and the processes giving chemical“texture” to the oceanic carbonic acid system componentsexert a primary control, along with temperature, pressure andsalinity, on the saturation state of seawater with respect tocarbonate minerals.

2. Major Marine Carbonate Minerals

2.1. General ConsiderationsIn this section, we provide a basic overview of past and

current work on the physical chemistry of carbonate mineralsrelevant to their reactions in modern marine environments.An accurate understanding of the relationship betweenseawater saturation state and the distribution of reef, algal,and associated carbonates in modern environments is alsocritical in evaluating the implications of rising sea surfacetemperatures and increasing acidity associated with anthro-pogenic CO2 fluxes. Over the past half century, extensivelaboratory and experimental work on carbonate minerals hasyielded important insights documented in a voluminousliterature. This section can by no means be encyclopedic butdoes attempt a selective review of seminal papers that havedefined major benchmarks in understanding.

It is critical to recognize that the vast majority ofsedimentary carbonate minerals deposited in modern marineenvironments isbiogenic in origin; that is, these mineralsare actively precipitated from seawater by organisms to formskeletal hard parts. A much smaller fraction of the totalsedimentary burden of carbonate minerals is formed abio-

genically, typically either as heterogeneously precipitatedprimary marine cements or through postdepositional, diage-netic reactions. Regardless of the organism responsible fortheir formation, biogenic skeletal carbonates bear distinctdifferences from their abiogenic counterparts in terms ofbasic physical properties, composition, structure, and reactiv-ity. In many ways, biogenic carbonates form an entirelydifferent class of materials, sufficiently complex and het-erogeneous that intrinsic relationships between physical andchemical properties depart fundamentally from those ofinorganic phases.

2.2. Physical and Chemical Properties

2.2.1. General Considerations

The crystal structure of naturally occurring carbonates, asa group, is relatively simple, particularly when comparedwith oxides, silicates, and other sedimentary minerals.Despite their structural simplicity, the behavior of theseminerals in seawater is complex. This complexity primarilyreflects the fundamental role of chemical kinetics in thissystem. As a diagnostic illustration, a simple reaction pathcalculation predicts the formation of ordered dolomite(CaMg(CO3)2) in normal surface seawater. However, despiteextensive supersaturation, unambiguous observations ofdirect dolomite precipitation from normal seawater areessentially nonexistent; aragonite and magnesian calcite arethe phases that actually form, at least via abiotic reactions.Thus great care must be taken in the use of simplethermodynamic relationships as predictor functions in sea-water systems where carbonate mineral distributions aredominated by metastable phases and where reactions maybe governed by poorly understood surface phases whoseproperties differ from the underlying bulk mineral. Asdiscussed later on, our current understanding of the kineticmechanisms that control precipitation and dissolution of thesephases is in some cases quite limited and requires both directobservations and an adequate theoretical framework. Webegin with a basic review of the unique physical andchemical properties of marine carbonate minerals.

The most important carbonate minerals in seawater reac-tions are the CaCO3 polymorphs aragonite and calcite.Marine calcite may also contain variable amounts (10 to>30mol %) of MgCO3 in solid solution, generically termedmagnesian calcite. The solubilities of some carbonate miner-als are given in Table 1. Aragonite is both denser and moresoluble than calcite. These two phases also differ in theirtendency to accept divalent cations other than Ca2+ in solidsolution. Thegeneralpattern of coprecipitation is governedby the following rule. Cations such as Sr2+ and Ba2+ whoseionic radii exceed that of Ca2+ (1.00 Å, assuming 6-foldcoordination with oxygen3) are more frequently accom-

Table 1. Stabilities of Common Carbonate Minerals under STP Conditionsa

phase formulaK′sp(stoich)

((mol2 kg-2) × 10-7) pK

aragonite CaCO3 6.657 8.30155

calcite CaCO3 4.397 8.48155

magnesite MgCO3 8.049

siderite FeCO3 10.529

rhodochrosite MnCO3 10.089

dolomite CaMg(CO3)2 18.159

kutnahorite CaMn(CO3)2 21.81( 0.0710,39

a K′sp(stoich)is the stoichiometric solubility product in seawater. pK values are thermodynamic-log Ksp.

344 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 4: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

modated in the orthorhombic aragonite structure, in whichcations are in 9-fold coordination to carbonate oxygens.Conversely, smaller cations such as Mg2+ are found morefrequently in the hexagonal calcite structure, in which thecation is in octahedral (6-fold) coordination with oxygen,each of which belongs to a distinct carbonate (Figure 1).These coprecipitation trends are at best qualitative statements,and observed distributions have shown poor fidelity withattempts to define partitioning on purely thermodynamicgrounds.4 In general, it is difficult to make robust statementsregarding the incorporation of “trace” elements into carbon-ates: careful studies of partition coefficients show that theseterms are in many cases dependent on both precipitation rateand tracer concentration, thus obviating the expectation of aunique Doerner-Hoskins5 proportionality between surfacesolid and solution tracer/carrier ratios. This issue is discussedin detail in Morse and Bender.6 In complex multicomponentsolutions such as seawater, the potential for interactionsbetweentracers also introduces significant additional com-

plexity in understanding or modeling partitioning behavior.Phases other than aragonite and magnesian calcite are of

far less abundance and importance. Vaterite (µ-CaCO3) ismetastable with respect to calcite and aragonite and limitedto minor biogenic occurrences; hydrated CaCO3 phases (e.g.,ikaite) are found only in very low temperature, high-pressureenvironments. Solid solutions and double carbonates arediscussed below.

It should be noted that temperature and pressure varyenough in the ocean to have a significant influence oncarbonate mineral solubility and equilibrium constants forthe CO2-carbonic acid system. Extensive discussion of theinfluences ofT andP can be found in refs 1 and 2.

2.2.2. R3hc Carbonates: Calcite, Magnesian Calcite, andRelated Phases

Numerous books and monographs already contain exten-sive reviews of the chemical and physical properties ofcalcite,1,16 and this discussion will thus serve only as arelatively brief summary of the issues relevant to marinesystems. The structure of theR3hc carbonates is best seenusing the hexagonal unit cell, with calcite as the example. Ifthec-axis of the calcite unit cell is oriented vertically, calciumcations form horizontal layers separated by planar CO3

groups (Figure 1). The distribution of carbonate solidsolutions within this structure primarily reflect differencesin cation size.13 Complete miscibility exists between MgCO3,FeCO3, and MnCO3 end members, whose differences in ionicradius aree0.11 Å in octahedral coordination. CaCO3-MgCO3 miscibility is discussed in detail below. Magnesite(MgCO3) is virtually unknown in open marine environments,and its properties are primarily of interest in terms of theirrelationship to the other magnesium-bearing phases. Divalentiron (siderite, FeCO3) is relatively rare in normal marineenvironments, because Fe(II) is typically first sequesteredas sulfidic phases during microbial sulfate reduction. Sideriteis thus more typically found in non-marine reducing environ-ments. In contrast, mixed calcium-divalent manganesecarbonates are potentially important in reducing environ-ments, and interactions of Mn2+ with carbonate surfaces arereviewed in more detail below.

The influence of magnesium on the properties andbehavior of calcite in seawater is a long-standing andfundamental problem. Several key issues hinge on the precisenature of this interaction: (1) the accurate expression ofcalcite solubility as a function of magnesium concentrationin the lattice, (2) how the properties and behavior of abioticmagnesian calcite differ from those of their biotic counter-parts, and (3) how these terms control the rate of dissolutionand precipitation reactions. We discuss the last issueseparately in the sections on precipitation and dissolution;calcite solubility and factors that contribute to measuredvariations thereof are discussed in detail below.

Demonstration of equilibrium typically requires approachfrom conditions of both over- and undersaturation. Inseawater, however, significant coprecipitation of MgCO3

generally results in a (abiotic) magnesian calcite (∼11 to 14mol % MgCO3); abiotic precipitation of a pure, Mg-freecalcite is not observed. This is expected: simple Gibbs-Duhem relations demand that the chemical potential ofMgCO3 in the solid (magnesian calcite) and aqueous(seawater) phase be equivalent at equilibrium. Uncertaintyin the solubility of magnesian calcites has been a source ofhistorical controversy, in part animated by basic disagree-ments as to the nature of the equilibrium condition with

Figure 1. Carbonate mineral unit cells (Ca in green, C in black,O in red): (A) calcite hexagonal unit cell, viewed down thec-axis,showing coordination of metal atoms by oxygens belonging todifferent carbonate groups;11,12 (B, left) relationship of hexagonaland rhombohedral unit cells (green and blue outlines, respectively,c-axis vertical) to cleavage rhombohedron (red outline, not a trueunit cell13,15); (B, right) hexagonal unit cell, same orientation;11,12

(C) orthorhombic unit cell of aragonite,c-axis vertical, showingstaggered orientation of trigonal carbonate groups.11,14 Crystal-lographic models were generated using PLATON software.11

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 345

Page 5: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

respect to this phase. Early attempts to measure magnesiancalcite solubility adopted a somewhat controversial approach.For example, in free drift, pure water-fixed pCO2 dissolutionexperiments, Plummer and Mackenzie17 attempted to cir-cumvent lengthy equilibration times by extrapolating solutionactivities to infinite time (t-1/2 f 0).

This “kinetic” approach to true saturation is sensitive tothe onset of incongruent reaction legs and, in general, hasbeen regarded as somewhat controversial.18 There is also thepotential for large variability in the reactivity of biogenicphases, discussed below.

Formation of magnesian calcite can be expressed as

wherex is the mole fraction of MgCO3 in solid solution.One means of expressing the solubility of magnesian calciteis by writing the thermodynamic equilibrium constant forthis reaction (K),

wherex is taken from eq 2.1 andai is the activity of speciesi. The problem with this expression, however, is that itpredicts relationships between the composition of the solutionand the solid phase (x) that may not, in fact, exist. This canbe seen by reorganizing eq 2.2 in terms of the cationconcentration ratio, whereγi is the total ion activity coef-ficient of speciesi,

Putting aside for the moment the issue of activity coef-ficients, for a given calcium and carbonate ion concentration,we are still free to vary the left-hand cation ratio by changingmagnesium concentration. For the equality to be preserved,however, the right-hand side of eq 2.3 would have to varysimultaneously in some complex fashion: the action ofaqueous magnesium concentration (or activity) has noexplicit representation here. More to the point, solid-phasecomposition (x) could be varied at no apparent expense tothe left-hand side, implying an arbitrary range of possiblesolid compositions for a given cation ratio. This freedom isinconsistent with what is actually observed in seawaterexperiments (see below), where precipitation of calcite at agiven Mg/Ca seawater ratio yields a unique solid-phasecomposition. Moreover the observed relationship of this solidwith the solution cation ratio is linear inx.

In a theoretical treatment, Thorstenson and Plummer19

argued that if magnesian calcite reacted as a solid offixedcomposition, then the Gibbs-Duhem equality of chemicalpotentials of individual components (i.e., CaCO3, MgCO3)in solid versus aqueous phases no longer has meaning. Inthe case of a congruently dissolving magnesian calcite, thisentails a loss of a degree of freedom because of the kineticrestriction of mass transfers of components with respect tothe solid. Thorstenson and Plummer invoked the term“stoichiometric saturation” to describe this condition. Itshould be emphasized that because of this path constraint,stoichiometric saturation states would form a superset of thetrue (stable or metastable) equilibrium condition. In CaCO3-MgCO3 solid solutions, the primary point of their argument

is that mere equality of the ion activity products (IAP) andKeq (eq 2.2) may demonstrate stoichiometric saturation (µCa(1-

x)Mg(x)CO3(aq) ) µCa(1-x)Mg(x)CO3(s)) but is an otherwise insufficientdemonstration of equilibrium (µMgCO3(aq) ) µMgCO3(s), µCaCO3(aq)

) µCaCO3(s)). Thorstenson and Plummer applied this thinkingto the earlier dissolution data of Plummer and Mackenzie17

and concluded that true equilibrium, or for that matterstoichiometric saturation, was probably not achieved forcalcites with substantial mol % MgCO3: either conditionwould necessitate a solution Mg/Ca ratio or an IAPcalcite

atypical of natural systems. They thus concluded thatthermodynamic equilibrium with respect to magnesian calciteexists only for solids with less than∼5 mol % MgCO3.Arguments regarding stoichiometric saturation with respectto the magnesian calcites have been debated extensively.20-23

Morse and Mucci addressed the issue of magnesian calcitesolubility directly in a series of key experiments, taking adistinctly different approach from that described above(description of this work will also overlap with that ofprecipitation kinetics). Morse et al.7 measured the solubilityof pure calcite in contact with seawater for extended periods(from 5 days to 1 month). Their measured apparent constantfor calcite (K′sp ) 4.39× 10-7 mol2 kg-2, see Table 1) wasrelatively consistent with other contemporary values, al-though less (∼20%) than that computed from the thermo-dynamic solubility product for calcite and total activitycoefficients for calcium and carbonate ion, a discrepancy theyattributed to the formation of a magnesian calcite surfacephase. This latter hypothesis was investigated in extensiveexperimental work,24-26 and because of their bearing on howsaturation develops between calcite and Mg-bearing fluids,we enumerate the principal findings of these papers in moredetail below.

1. The extent of Mg uptake is a function of thesolution concentrationratio [Mg/Ca]soln, not theabsolute concentrations of each component. Usingdata from pH-stat precipitation runs, Mucci andMorse24,25 argued that the behavior could be bestunderstood in the context of an “exchange equi-librium” between the crystal surface and the solu-tion and between surface and the underlying crystalemphasis added. At low solution ratios ([Mg/Ca]soln

< ∼7.5), surface adsorption of Mg is favored overCa, such that [Mg/Ca]surf > [Mg/Ca]soln. Under theseconditions, sorbed Mg has the greatest influenceon the net incorporation of Mg into the overgrowthitself, and the ratio of [Mg/Ca]solid/[Mg/Ca]soln isnegatively correlated to the [Mg/Ca]soln. The amountof Mg incorporated within the lattice in this regionis significant but small, limited by the Mg contentof the solution.24

2. As solution ratios increase ([Mg/Ca]soln >∼7.5), the ratio of [Mg/Ca]solid to [Mg/Ca]soln

approaches the constant value of a classical homo-geneous distribution coefficient (DMg). However,Mucci and Morse24 argued that this observation wasinconsistent with a simple Langmuir adsorptionisotherm, because progressive saturation of surfacesites should give rise to a variableDMg withincreasing [Mg/Ca]soln. They thus concluded that afeedback control is established at the growingcrystal’s interface, in which the formation ofmagnesian calcites having preferential affinity formagnesium establishes a dynamic equilibrium be-

Ca(1-x)Mg(x)CO3 ) (1 - x)Ca2+ + xMg2+ + CO32-

(2.1)

Kmc(x) ) aCa2+1-xaMg2+

xaCO32- (2.2)

[Mg2+]

[Ca2+])

γCa2+

γMg2+ ( Kmc(x)

aCa2+aCO32-)1/x

(2.3)

346 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 6: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

tween the solution, the surface, and the nascentovergrowth. Overgrowth composition may ad-ditionally reflect crystallographic control.20

3. The above data link a given [Mg/Ca]soln witha corresponding [Mg/Ca]solid. In free drift experi-ments, Mucci and Morse25 subsequently derivedstoichiometric solubility constants from seawatersolutions of variable [Mg/Ca]soln equilibrated withpure calcite. On the basis of these results, theyconcluded that MgCO3 coprecipitation doesnotchange calcite solubility, as computed from the ionactivity product for CaCO3; if ion pairing isaccurately represented, calcite saturation can beadequately described by the CaCO3 ion molalproduct alone (K′c ) [Ca2+]|eq[CO3

2-]|eq); cf. eq2.2). Furthermore, they introduced the notion thata “solubility-controlling” phase of invariant com-position satisfied the exchange equilibrium identi-fied above. Furthermore, Bischoff et al.,27 incomputing solid phase activity coefficients, con-cluded that magnesian calcites forming as over-growths in Mucci and Morse26 were indeed equi-librium phases. Most importantly, magnesian calcitesolubilities, recalculated from these data accordingto eq 2.2, are substantially less than those measuredcompared with the biogenic phases measured previ-ously (see below). These relationships are sum-marized in Figure 2.

In general, the solubility versus composition data show asolubility minimum at approximately 2 mol % MgCO3, withincreasing solubility thereafter. Although outstanding issuesremain, the current consensus is that the solubility versuscomposition data (Figure 2) can be divided into three basictrends. The first trend (curve A) incorporates well-crystal-lized, inorganic solids from the work of Mucci and Morse26

and Bischoff et al.28 This curve defines the lowest solubilitiesfor magnesian calcites, and the intercept atx ) 0 of the fitted

curve is very close to the accepted value for pure calcite(pK ) 8.48 at 25 °C). Although these data representequilibrium approaches from both over- and undersaturatedconditions and include material formed under both roomtemperature and hydrothermal conditions, the data set’soverall variability is remarkably small. An important obser-vation is that the increase in solubility withx is sufficientlysmall that abiotically precipitated magnesian calcites withgreater than or equal to∼20 mol % MgCO3 are all lesssoluble than aragonite (see below).

The second trend (curve B) is a “best fit” through thebiogenic, skeletal magnesian calcites.18,27 These data showsignificantly higher solubilities (∼0.15 differences in pK, orabout∼1.1 kcal mol-1), although the overall trend inx issimilar to that of the abiotic phases. The intercept at zeroMgCO3 also places this phase very close to aragonite and a20 mol % MgCO3 synthetic phase in solubility. The originof this increase in free energy is unclear but may includesubstantial contributions from excess water and impuritydefects (sulfate, hydroxyl, and bicarbonate); these contribu-tions would be present independent of MgCO3 content. Inaddition, it is well-documented that positional disorder ofthe carbonate anion (rotation out of the basal plane) isassociated with both abiotic and biogenically formed mag-nesian calcites,32 a trend that is also reflected in unit cellc/aaxial ratios. In general, the variation in solubility of biogeniccalcites with 12e x e 15 mol % MgCO3 is also greaterthan the corresponding variation in the synthetic phases. Animportant conclusion from these findings is that magnesiumcontent is not an accurate predictor of solubility for biogenicphases. Reactivities of biogenic phases are thus most likelydominated by characteristics introduced directly or indirectlyby the organism itself.

Most biogenic magnesian calcites from curve B in the 12-15 mol % MgCO3 range are slightly more soluble thanaragonite. However, these solubilities are substantially lessthan those recovered in the dissolution experiments ofbiogenic magnesian calcite by Plummer and Mackenzie,17

subsequently revised by Thorstenson and Plummer,23 inwhich 12-15 mol % MgCO3 magnesian calcites exceedaragonite solubilities by a factor of∼5. These latter data(curve C of Figure 2) may likely reflect the influence ofkinetic, rather than thermodynamic, factors, including theretention of reactive surface particles after minimal samplecleaning and lack of annealing but may also reflect reactivityin nature. These issues are also discussed in Bischoff et al.33

2.2.3. R3h Carbonates: Dolomite and Other DoubleCarbonates

Dolomite CaMg(CO3)2 is the most important doublecarbonate. It is found primarily in marginal, restrictedregimes that suffer extremes in terms of temperature,saturation state, or salinity or that otherwise differ fromnormal open marine environments. There are two oft-citedtruisms involving this mineral. First, it is a significantcomponent of ancient sedimentary carbonate rocks, incontrast to its meager present-day distribution. Second, therehas been little success in synthesizing dolomite at lowtemperatures in the laboratory. These two observations arethe basis of the oft-cited “dolomite problem”. However, thisperception may reflect inadequate recognition of (1) differ-ences in laboratory versus geologic time scales and (2) strongenvironmental controls over carbonate mineral distributions.In general, we still require a better understanding of how

Figure 2. Solubility of magnesian calcite computed from eq 2.2.Data are fromsynthetic(gray solid circles, Busenberg and Plum-mer;29 open circles, Bischoff et al.;27 black solid circles, Mucci andMorse;25 open squares, Lafon30), natural inorganic(crosses, Busen-berg and Plummer29), and biogenic phases (open diamonds,Busenberg and Plummer;29 closed triangles, Bischoff et al.;27 closeddiamonds, Walter and Morse;18 solid squares, Plummer andMackenzie17 recalculated by Thorstenson and Plummer19). TrendsA, B, and C are discussed in text. Adapted from ref 31 withpermission. Figure was redrafted with data added. Copyright 1991Mineralogical Society of America.

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 347

Page 7: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

ordered mineral phases form at low temperature and, inparticular, how microbial communities may mediate thesereactions.34,35

Determinations of the calcite-dolomite solvus at hightemperatures show complete miscibility between pure calciteand carbonates of dolomite composition only above tem-peratures of 1075°C.36 Composition and cation orderingdistinguish ideal dolomite from magnesian calcite. Phasesthat deviate from dolomite’s cation stoichiometry (e.g., Ca/Mg > 1) and show enlarged unit cells and diffuse diffractionpeaks are collectively termed “protodolomites”. Graf andGoldsmith first used this label to refer to both synthetic andnaturally occurring materials that would presumably react,given sufficient time, to form ideal, fully ordered dolomite.37

Ordered dolomite is essentially a calcite structure in whichevery other cation layer is fully occupied by magnesiumatoms. This results in a symmetry reduction (R3hc f R3h)through loss of calcite’sc-glide axis. In contrast, magnesiumatoms in a magnesian calcite are not segregated to their owncation plane but substitute for calcium in a completelyrandom manner. Sedimentary dolomites are often poorlyordered and calcium-rich, with increasing disorder detectableas a progressive weakening in the superstructure reflections(101), (015), and (021).38 Other metal ions may alsosubstitute for magnesium, notably iron and manganese;ferroan dolomite (ankerite) can accommodate a maximumof ∼70% CaFe(CO3)2. The manganoan equivalent of dolo-mite, lying on the CaCO3-MnCO3 join, is termed kutnahorite(CaMn(CO3)2).39

We know of no solubility determinations for dolomite inseawater. Because of the large effects of ordering andcomposition, the uncertainties surrounding dolomite stabilityare quite large. Determinations of enthalpies of formation(∆H°f) of ideal, ordered dolomite vary by several kilojoules.Because the contribution ofT∆S at Earth surface tempera-tures is small (with good agreement on entropy determina-tions), these variations in measured enthalpy yield corre-sponding uncertainties in free energy. An early attempt tomeasure dolomite solubility by a similar approach to thatdescribed above for the magnesian calcites (extrapolation toinfinite time) shows a very low value of (pK25°C ) 19.740);current accepted pK values range from 18.1 to 17.3. Theeffect of ordering and composition on dolomite stability hasbeen investigated extensively by Navrotsky and co-workers.41

Solution calorimetry42 showed that enthalpies become stronglyendothermic with increasing calcium enrichment; similardeterminations for magnesian calcites33 show negative en-thalpies by comparison, so clearly calcite accommodatessubstantial lattice magnesium with far less loss of stabilitythan dolomite accommodates calcium.

2.2.4. Orthorhombic Carbonates: Aragonite

Aragonite constitutes a major phase of marine sediments,occurring both as marine cement and as the principal skeletalcomponent of many marine taxa. Sr2+ and Ba2+ are importantcoprecipitating cations for Ca2+ in 9-fold coordination inaragonite. The solubility product for aragonite in pure(nonseawater) solutions is relatively well-known, with a pKa

at 25 °C ranging from 8.34( 0.0243 to 8.28( 0.02.44 Inseawater, determinations of the stoichiometric constant (pK′a, note thatK′a has units of mol2 kg-2) range from 6.09(0.0345 to 6.18 ( 0.02.7 The substantially lower solubilityfound for aragonite by Morse et al.7 reflects the exceedinglylong equilibration time (2 months) used in their seawater

determinations, as previously published determinations em-ployed shorter (hours to days) periods. This measured valueis 22% less than that predicted for aragonite solubility fromthermodynamic solubility products (pKa) and total ion activitycoefficients in seawater; however, unlike the magnesiancalcites, no recognizable difference exists between solubilitiesobtained from synthetic versus biogenic materials. It iscritical to recognize the dependence of the apparent solubilityproduct of aragonite on equilibration time (Figure 3). Thisresult is observed consistently, is apparently reversible, andindicates unsolved complexities in the behavior of aragonitein seawater. The observed reduction in solubility may reflectthe formation of a surface phase of (as yet) unknowncomposition and structure, although aragonite dissolutionafter extended exposure to seawater cannot obviously beginuntil seawater is undersaturated with respect to this putativesolubility-controlling phase. A time-sensitive surface phaseis also consistent with experimental results of Tribble andMackenzie,46 in which the composition of neoformed over-growths are not constant but tend to alter with time as thenascent phase progressively excludes Mg from the lattice.Thus the solid-phase composition is both a function of time,and in the case of a calcite overgrowth, depth from thesubstrate. This problem also has direct bearing on the issueof “kinetic solubility” and the response of seawater saturationstates to pCO2 changes as discussed later in this paper.

3. Carbonate Mineral Precipitation andDissolution Kinetics Relevant to Seawater

3.1. General OverviewA successful approach to the general problem of calcite

and aragonite reaction kinetics in seawater solutions mustsolve several difficult problems. These derive both fromgeneral issues related to the kinetics of mineral-solutioninteraction and from the specific properties of seawater andcarbonate minerals as reactants. We give a brief overviewof general concepts first and dwell on the latter in specificsubsequent sections.

Although seawater chemistry is covered extensivelyelsewhere in this issue, before continuing we should makesome brief remarks concerning the carbonic acid system inseawater and its relationship to carbonate mineral reactionkinetics. First, constraint of the carbon speciation in seawaterrequires a minimum of two measured parameters: pH, pCO2,dissolved inorganic carbon (DIC), or total alkalinity (in

Figure 3. Trend of equilibration time and apparent solubility ofaragonite. Modified from ref 7, Copyright 1980, with permissionfrom Elsevier.

348 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 8: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

addition to salinity or ionic strength, temperature, pressure,etc.). Dissolution or precipitation must at some point in thereaction sequence involve carbonate ion, and thus carbonateion concentration and activity are clearly key quantities. Thelog activity of carbonate ion can be computed as the sum oflog K2 + log pCO2 + 2pH.1 However, this computed activityis relevant for the bulk solution; at the mineral surface itself,the distribution of surface and near-surface species reflectsboth equilibrium hydration and hydrolysis reactions, as wellas the kinetics of the dissolving or precipitating mineral itself,and thus may differ significantly from the bulk solution.

3.2. Mineral Reactivity and the Mineral −SolutionInterface: The Link Between ThermodynamicDriving Force and the Overall Rate

3.2.1. General Rate Equations

A governing principle in the chemical kinetics of precipi-tation or dissolution is that the overall process can be dividedinto a sequence of discrete reaction steps: (1) transport ofmaterial to the mineral surface from the bulk solution throughsome sort of boundary layer environment, (2) adsorption ofreactive solutes to the surface itself, (3) surface diffusion,essentially a random walk down a concentration gradient tothe reactive site, and (4) once the reactant arrives at this site,the reaction itself may proceed through a series of individualsteps entailing bond formation or cleavage, ionic exchangewith the solid, acquisition or loss of solvent water, and othersteps. In the case of dissolution, the products of the reactionare original lattice components as solutes, which may thenfollow the reverse of diffusion and transport steps 3 through1, resulting in their final departure from the surface to thebulk solution. Detailed treatment of these steps is availableelsewhere.47 Our key point here is that the slowest step inthis series is limiting and thus determines the rate of theoverall process. For the majority of carbonate mineralreactions involving seawater, the rate of integration ordisintegration of ionic components at the reactive site itselfis the slowest step in this chain. This limitation reflects therelative insolubility of most carbonate minerals, as well asthe confined range of disequilibrium (extent of over- andundersaturation) in marine environments, at least within thewater column. Thus in general, carbonate precipitation anddissolution rates in seawater aresurface-controlled.

In order to understand variation in the rates of mineralsurface processes, a basis must exist to evaluate reactionmechanisms and their relationship to primary environmentalvariables. The complexity of seawater poses a fundamentalchallenge in this requirement. A substantial number ofobservations now exist of the overall carbonate mineralreaction rate’s dependence on saturation state, temperature,salinity, and other variables. However, these results have beenunderstood primarily on an empirical basis, and conclusionsregarding mechanism are often speculative. This limitationis slowly receding, however, with the integration of directobservations of the reacting mineral surface, discussed below.We preface this with a review of current rate equations.

Thermodynamic equilibrium establishes the limiting con-dition at which the overall reaction rate must be zero. Asdiscussed at length in Lasaga47 and other texts, the relation-ship of the equilibrium point and reaction rate is oftenillustrated through the application of detailed balancing. Thisapproach implicitly recognizes the principle of microscopicreversibility, in which the maintenance of equilibrium

provides that the forward reaction path followed by amolecular assembly be the reverse equivalent of the back-ward reaction. This condition allows the ratio of rateconstants for these reactions to be related to the equilibriumconstant. For example, if the reversible reaction of reactantR and product P are represented as the forward- and back-reactions,

the forward and backward rates are proportional to theconcentrations of R and P, respectively. At equilibrium, theserates will be equivalent:k+[R]eq ) k-[P]eq. Under thisequality the ratio of rate constants is proportional to theequilibrium constant,

whereγ is the coefficient relating concentration and activityterms. This oft-described relationship provides a basic linkbetween equilibrium thermodynamics and chemical kinetics.Lasaga47 used detailed balancing to derive the generalreaction for calcite dissolution,

Here r is the rate,A/V is the ratio of total mineral surfacearea to solution volume, andKsp

n andΩn are the solubilityproduct and saturation ratio (Ω ≡ aCa2+aCO3

2-/Ksp) for calcite,respectively, both raised to the power of the parametern.Equation 3.3 is identical to that derived by Sjo¨berg fromexperimental work with calcite in 0.7 M KCl solutions (pH8-10), in whichn assumed a value of1/2.48 However, thisequation does not accommodate any dependency that thedissolution rate may have on the composition of the solution,other than that explicitly provided by the saturation stateterm. If it is assumed to hold over large changes in saturationstate, it also makes an implicit assumption with respect tomechanistic control (see discussion below). In contrast,empirical expressions generally employed for the dissolutionof carbonates represent the overall rate as49

An equivalent expression for precipitation can be formedby a simple change of sign. As discussed in more detail inMorse and Arvidson,50 eqs 3.3 and 3.4 clearly differ,particularly in terms of the function of the parametern. Ineq 3.4,n is the order of the overall reaction; in eq 3.3,nacts on the product of activities and thus operates identicallyon the activity of calcium and carbonate ion. The latterrelationship reflects simple reaction stoichiometry in eq 3.3,where net steady-state dissolution must liberate equalamounts of oppositely charged lattice ions.

More generally, there is a critical distinction between“linear” rates, such as the following,

where the- and+ signs apply to dissolution and precipita-tion, respectively, andn ) 1/2, n ) 1, etc.,51,52 versus the

R \k+

k-P (3.1)

k+

k-)

[P]eq

[R]eq

)γR

γpKeq (3.2)

-r ) k-AV

Kspn(1 - Ωn) (3.3)

-r ) kAV

(1 - Ω)n (3.4)

(r ) (k((Ωn - 1) (3.5)

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 349

Page 9: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

commonly used expression in eq 3.4. Truncation of a Taylorseries expansion for the free energy function in eq 3.5 (recallthatΩ ) e∆Gr/(RT)) shows that as∆Gr approaches 0, e∆Gr/(RT)

≈ ∆Gr/(RT) + 1, and the rate (written here as dissolution)thus becomes linear in free energy:

However, even close to equilibrium, our conventionalexpression, eq 3.4, for carbonate reactions remains nonlinear(assumingn * 1), and more generally

where|∆Gr| denotes the absolute value of free energy, thenegative sign applies to dissolution, and the positive appliesto precipitation.53 This is not a trivial point concerning thefitting of rate data, but a fundamental insight into the roleof reaction mechanism. Lasaga elegantly emphasizes thispoint in terms of irreversible thermodynamics: expressionssuch as eqs 3.4 and 3.7 are inconsistent with the guaranteeof microreversibility implied by eqs 3.1 and 3.2. Why?Becausedislocations are not equilibrium defects;47 if growthor dissolution of carbonate surfaces is limited at low super-or undersaturations, respectively, by nonequilibrium surfacedefect distributions, then these surfaces cannot be consideredto be at equilibrium even for∆Gr ≈ 0. Nielsen also suggestedthat rate equations having the general form of a concentrationdifferenceraised to some power (e.g.,r ) k(c - s)2) likelyreflect the nonlinear dependencies inherent in spiral disloca-tion-controlled crystal growth.54 Thus, in the case of carbon-ates, both the general form of rate equations (eqs 3.4 and3.7) and experimental evidence (see below) point to thefundamental role of defect control in both dissolution andgrowth kinetics. Limited attempts have also been made toexpress explicitly the rate’s control by defect density.55

Strained calcites have been used to show that the influenceof dislocations could be accommodated by differentiatingbetween perfect surface sites and dislocation sites, eachcharacterized by distinct energy distributions and corre-sponding rate constants. Although reaction rate does not scalelinearly with defect density, Schott et al.56 found thatdislocation densities greater than 107-108 cm-3 result in anincrease in dissolution rates under near-equilibrium condi-tions.

Despite this evidence of nonlinear controls on reaction rate,many practitioners incorrectly use simple carbonate ionconcentration difference terms to drive reaction rate expres-sions. There are additional reasons for application of first-order linear kinetics for carbonate precipitation and disso-lution in seawater to be incorrect in practice: in addition tothe potential variation of reaction mechanism with changingsaturation state, dissolved impurities may play a fundamentalrole. For example, there is substantial evidence that dissolvedmagnesium and phosphate inhibit both precipitation anddissolution reactions in seawater (discussed below). Impuri-ties may strongly influence the reaction rate itself via simplesorption but may also exert more complex, mechanisticcontrols, changing the nature of the rate’s relationship tosaturation state or controlling the mineralogy and composi-tion of the phase that nucleates on pre-existing crystalsurfaces. Surface interactions with impurities are not under-stood in detail, although recent experiments suggest possible

reasons for this inhibition to couple strongly with saturationstate. In addition, simple first-order expressions are virtuallycertain to fail in pore water environments where theconstancy of major ion ratios is not maintained.

A more general rate equation can be used to represent theaction of catalyst and inhibitor species, and Lasaga57 hassuggested the following form to accommodate these terms:

Here the Arrhenius temperature dependence can be shownexplicitly as

wherek0 is the intensive, temperature-independent prefactor,Ea is activation energy, andA is reactiVe surface area. Thisexpression, applicable to both precipitation and dissolution,provides a means of calibrating the effect of ionic strength,g(I), independent of that provided by the use of activitycoefficients, as well asai, the action ofN inhibitors, catalysts,proton activity, etc. The termf(∆Gr) is the free energy termdescribing the rate’s functional relationship to disequilibrium.If we neglect the inhibition and ionic strength terms, replacef(∆Gr) with a saturation state function, administer signs suchthat ∆Gr corresponds to the dissolution reaction, and allowthe rate constantk to collect the (reactive surface) area andvolume terms, we effectively recover our conventionalexpression (eq 3.4),

With increasing undersaturation,∆Gr f -∞, f(∆Gr) f 1,and the dissolution rate approaches a constant valuek, thedissolution plateau. With increasing oversaturation, that is,exp(∆Gr/(RT)) - 1, the precipitation rate increases expo-nentially; however, the precipitation or crystal growth rateobviously cannot increase infinitely. As sufficient differencein chemical potential develops, new heterogeneous andultimately homogeneous crystal nuclei will appear whoseminimum size reflects the system’s critical radius (discussedin terms of Gibbs-Thompson relations in nucleation below).Note that although the form of the equations described abovepermits application to both growth and dissolution, differentvalues ofk andn would obviously be used for each.

3.2.2. Relationship to Mineral Surface ProcessesAt the core of the existing theoretical framework for

surface-controlled reaction rates is the concept of surfacefree energy distribution. Because precipitation or dissolutionrates bear an exponential energy dependence (see previousrate equations), heterogeneity in this distribution potentiallygives rise to large variations in rate over the mineral surface.This heterogeneity reflects the physical distribution ofdislocations in the crystal (plane, line, and point defects, seeFigure 4). Seminal work by Burton et al.58 recognized thecentral role of screw dislocations in terms of providing acontinuous source of atomic surface steps during crystalgrowth. More recently, direct surface observations of thecleavage surfaces of rhombohedral carbonates59-61 haveshown that step movement can be related to the overall ratemeasured in bulk solution (see also Figure 5). The symmetryrelations of kink and step geometry in calcite also controls

r ) k(1 - exp(∆Gr/(RT))) f -n∆Gr

RT(3.6)

r ) -k[|∆Gr|RT ]n

(3.7)

r ) kAg(I) ∏i

N

aini f(∆Gr) (3.8)

k ) k0 eEa/(RT) (3.9)

r ) -k(1 - exp(∆Gr/(RT)))n (3.10)

350 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 10: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

site-specific uptake of impurities. The control of step velocity,the manner by which steps propagate, and the overalldistribution of steps on the surface (step density or spacing)is thus the link between reaction mechanism and itsphenomenological expression as a material flux betweenmineral and solution.

Recognition that the speciation of surface sites may becorrelated with the reaction kinetics has led to the develop-ment of thesurface complexation model, developed andinitially applied to calcite dissolution by Van Cappellen etal.62 This model is entirely analogous to the system of relatingthe distribution of aqueous complexes and ion pairs byequilibrium mass and charge balance expressions. Thus theratio of surface complexes varies in concert with that ofdissolved carbonic acid species, establishing a relationshipbetween the surface concentrations and dissolved activities,and yielding potential insight into surface-controlled reactionkinetics. For example, Arakaki and Mucci63 derived a generalrate equation for calcite dissolution and precipitation insimple solutions through application of this model to diversedatasets. This general approach has also been campaignedas a general solution to the problem of variation in carbonatemineral reactivity, arguing that dissolution reaction kineticscan be described in the context of rate-controlling surfacecomplex distributions.64 Recent attempts have also been madeto integrate this model with mineral surface observations.65

However, under conditions where reactive transport controlbecomes important, uncertainties in evaluating the diffusionof carbonic acid species may make this approach problematic.In addition to providing no means for transport via surfacediffusion, this approach also contains no intrinsic descriptionof the variation in reactive site distribution (e.g., attachment/

detachment at kink sites and steps versus terraces). This issueis discussed later in the text, as well as in Morse andArvidson.50

Although the previous section has emphasized the impor-tance of understanding reaction mechanism and its relation-ship to saturation state, there is currently no over-archingtheoretical framework that would relate the form andparametrization of rate equations driven by seawater satura-tion state to surface reaction mechanisms. Most experimentalmeasurements of mineral reaction rates, for example, in alaboratory flow-through or batch reactor, are done throughevaluation of changes in the composition of the bulk solution.Changes in the mineral surface, and thus the link betweenrate and mechanism, are not observed directly. There is alsoa long-standing recognition of the problem of surfacereactivity distribution, described briefly above and discussedin more detail below in section 3.2.4. These limitations havegiven rise to an increasing interest in understanding therelationship between the macroscopic or “bulk” rate andmineral surface processes through direct observation, spon-sored initially by the advent of the atomic force microscope(AFM66,67). The AFM has been used extensively oncalcite,61,68-73 magnesite,74 and dolomite75 because of therelative ease of generating excellent cleavage surfaces. Inconcert with a suitable fluid cell, AFM permits direct,nanoscale observation of real-time precipitation and dissolu-tion reactions. In fluid, AFM provides angstrom-level verticalresolution at nanometer-scale lateral resolution. Rates aretypically measured by comparison of fixed area raster imagesas a function of time. This practice limits the measurementof very fast rate; conversely, measurement of very slow ratesis a function of the lifetime of the cantilever tip. However,there are problems associated with the extrapolation of AFMstep velocities to surface-area-specific rates measured byconventional powder methods: e.g., the relationship of stepvelocity to a surface-normal velocity is given by (Figure 5)

whereVn is the surface-normal velocity,h is step height (e.g.,0.3 nm for calcite),Vstep is the velocity of the step, andL isthe average step spacing. Although measurement of the stepheight is a trivial matter for the AFM, step spacing is adynamic property and thus requires care in acquisition. Inaddition, there is the more important problem of howlocalized measurements, conducted over small regions of themineral surface for brief duration, can be applied to largerspace and longer time scales. This problem is a fundamentalone and is discussed in more detail in sections 3.2.4 and3.2.5 below.

3.2.3. Treatment of Kinetic Inhibition

There has been recognition of the strong potential role ofreaction inhibitors on carbonate mineral reaction kineticssince the early work of metal inhibition of calcite dissolutionby Weyl76 and Terjesen et al.77 The problem with using pastresults to form an overall, integrated understanding of therole of inhibitors is the lack of a “standard model” forcomparing work in disparate bulk solution compositions,saturation states, temperatures, solid preparation, etc. Thusmuch of the existing results can be compared only on a

Figure 4. Schematic view of defects and specific sites on a crystalsurface: (1) terrace adatom; (2) terrace vacancy; (3) step vacancyor “double” kink site; (4) step adatom; (5) single kink site; (6)outcropping edge dislocation; (7) outcropping screw dislocation;(8) imminent hole.

Figure 5. Schematic representation of parameters relating crystalgrowth and dissolution rate with bulk rate: surface step height,h;kink spacing,x; interstep spacing (terrace width),L.

Vn ) hVstep

L(3.11)

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 351

Page 11: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

qualitative basis. Because of the nonlinear nature of bothgrowth and dissolution kinetics in carbonates in terms of freeenergy control discussed previously, comparison of the extentof inhibition at similar inhibitor concentrations but different∆Gr is neither practical nor desirable.

The theoretical treatment of inhibition on a surface derivesfrom a number of competing model approaches, for whichthe distinction hinges on the reversibility of the adsorptionstep. The Langmuir-Volmer model78 assumes that a revers-ible Langmuir adsorption isotherm prevails on the surface.Although this model is often employed as a means ofinterpreting inhibition, the reversibility requirement mayoften not have been demonstrated. In contrast, the “deadzone” model of Cabrera and Vermilyea79 assumes irreversibleadsorption of impurities on terraces or step edges. When thisscenario is applied in the context of crystal growth, advancingsteps are pinned and thus blocked from further advance asimpurities are encountered. Step pinning creates a “curved”step whose radius exceeds a critical value (see discussionon nucleation below). The reduction in step velocity leadsto a subsequent increase in saturation state of the solutionbecause of the reduced demand for lattice ions; whenoversaturation is sufficient, the pinned steps are able to breakthe impurity barrier and resume migration. As bulk impurityconcentrations increase, the “dead zone” of step pinning mayincrease proportionally.

If impurities bind irreversibly at appropriate surface sites,they may be physically incorporated within the solid duringcrystal growth. In this scenario of impurity incorporation,this produces a defect, resulting in a change in the strainfield of the crystal lattice and thus increasing the solubilityof solid crystal itself at this site (e.g., Davis et al.72). Thisaction results in a decrease in the driving force (∆µ betweensolid and solution), and step velocityV is decreased by afactor of 1- r*/ r, wherer* is the critical radius (see below),relative to the pure, impurity-free crystal growth system (V∞).In crystal dissolution, attachment of an impurity at a kinksite, rather than increasing strain, may conversely have theeffect of stabilizing the site against further detachment, thusinterfering with kink detachment dynamics and preventingproper step flow. As a whole, impurity interaction withcarbonate surfaces is a very complex topic, because itinvolves both the surface chemistry and the changes broughtabout by variations in solution composition; see Gutjahr etal.80 for a more detailed discussion.

3.2.4. Mineral Reaction Rates and Surface Area

The following discussion is an abbreviated version of theone presented in Lu¨ttge,81 and we refer the interested readerto this book chapter for a more detailed discussion. Fluid-mineral interactions, for example, dissolution and growthprocesses of carbonates and other minerals, can be observedin two principally different modes, either directly at thesolid-liquid interface orswhat has been the standardprocedure in the pastsindirectly through changes in thesolution chemistry. For a variety of reasons most laboratorystudies that were designed to investigate reaction kineticshave successfully focused on the solution chemistry and itschanges. By increasing the ratio of crystal surface area toreactor volume, even very slow rates become measurable.This is important because slow reaction rates on the orderof 10-8-10-16 mol m-2 s-1 are often characteristic formineral dissolution and growth reactions at Earth surface

temperatures. Analytically and experimentally, the indirectapproach is often easier because the result of many micro-scopic processes can be quantified at a larger scale overlonger periods of time. Thanks to the dedication of manyexperimentalists, we can build today on a large pool of ratedata and observations (e.g., White and Brantley82 andreferences therein; Morse and Arvidson50 and referencestherein).

The left-hand sides of eqs 3.3-3.7 are equivalent to the“bulk” rate, expressed as a function of free energy (∆G)implicitly or explicitly proportional to surface area (A). Thebulk rate constantk may also assume, in addition totemperature, additional dependencies related to solutioncomposition. Experimentally measured bulk rates have alsobeen expressed in terms of the formation of surface molecularcomplexes by surface reaction with solution species, appliedprimarily to dissolution.63,83-92 This approach highlights thechemical bonding properties of mineral surfaces, as well asthe role of adsorption-desorption steps on the crystaldissolution rate. The equilibrium distribution of surfacecomplexes has been used to predict the pH dependence ofthe rate. The molecular approach, therefore, has emphasizedthe role of surface speciationon mineral surfaces as afunction of solution composition and temperature. The modelthen tries to identify particular surface complexes whoseactivities limit the overall dissolution rate.87,92,93

All experimental studies that use the indirect approachhave one major problem in common: all rates are integratedover the entire mineral surface. Thus, these experimentscannot resolve the surface distribution of the integrated rate,the reaction mechanism cannot be observed directly, andsurface area in contact with the solution has to be quantifiedand rates must be normalized by thereleVant surface area.The quantification of the relevant surface area, however, hasproven to be a nontrivial problem and one of concern.

A challenge is already the definition of the “correct”surface area with which to normalize the bulk rates or fluxesmeasured experimentally or in nature. Authors such asHochella;94 Brantley et al.;95 Brantley and Mellott;96 Jeschkeand Dreybrodt,97 and most recently Lu¨ttge98 have discussedthis problem in detail. Unfortunately, this problem remainsstill largely unsolved and indicates that our understandingof the processes at the mineral-solid interface is not yetsufficiently advanced to provide a satisfying theoreticaldescription.

Independent of this fact, as long as mineral-fluid reactionsare surface-controlled, we need to normalize bulk reactionrates,R, by the surface area involved in the process to obtaina rate constant,k, in units of moles per square meter persecond. This rate constant is important because it is a criticalinput parameter for any kinetic model, and, therefore, aprerequisite for our predictive capabilities. Consequently, thequantitative treatment of surface area has been givenconsiderable attention in the geochemical literature.94,97-99

Theoretically there are at least three different quantities: (1)geometric area, (2) total or Brunauer-Emmett-Teller (BET)surface area, and (3) the “reactive” surface area. In thiscontext, we can only summarize the discussion. Currently,only so-called BET surface area100 is a robust measurement.Geometric surface area is merely an abstraction or estimate,and reactiVe surface area is problematic and more of aconceptual tool than a directly measurable quantity. Geo-metric surface area is usually calculated from the crystaldimensions and ignores surface roughness (cf, Anbeek101).

352 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 12: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

This calculation may also involve assumptions of an idealgeometric body, such as a cube, rhombohedron, or sphere.If the geometric surface area is calculated for a mineralpowder, it will depend on the way the average grain diameterof the grain population is computed. Because of the natureof grain size distributions, it is likely that an estimate ofgeometric surface area derived from a model underestimatesthe true value.

The term “total surface area” is often used synonymouslywith BET surface area, which is measured by gas-adsorptiontechniques developed in 1938 by Brunauer, Edward, andTeller.100 The extent to which the BET surface area measure-ment reflects the total surface area is a function of both thegas used and the structure of the solid surface. Unless thecrystal surface is atomically flat, that is, there is no surfaceroughness, the total/BET surface area is always larger thanthe geometric surface area. Consequently, the calculatedreaction rate constants in moles per square meters per secondare slower than they would be by normalization withgeometric surface area. This is discussed in detail by Lu¨ttge.98

These model calculations show that the differences betweendifferent types of surface area measurements may changesignificantly during a dissolution reaction due to the develop-ment of etch pits and other dissolution features.

In general, the surface area (however described) willchange over the course of a reaction. The extent to whichthe surface is self-modifying reflects the balance betweentwo competing processes: (1) a particle’s tendency tominimize its total surface energy and exposure to the solutionby reducing its area to volume ratio and (2) the mechanisticrestriction of possible paths by which the bulk free energycan be minimized, through material transfer to solutionthrough propagation of kinks, steps, and etch pits, which tendto increase the surface area.

The situation may be significantly complicated if we takeinto account that the mineral surface is usually not homo-geneously reactive, and thus certain areas of the crystalsurface react faster than other areas. This has been observedby scanning electron microscopy (SEM), AFM, and verticalscanning interferometry (VSI) studies, which often show theformation of etch pits, hillocks, and other features at crystalsurfaces during the reaction process. These features are oftenlinked to some microdamage within the crystal lattice (e.g.,microcracks) or lattice dislocations (e.g., line defects suchas screw dislocations or edge or point dislocations) that formhigh-energy sites at the crystal surface. The observation thatsome parts of the surface are significantly more reactivewhile others remain more passive has led to the idea ofreactive surface area.48,94,102Unfortunately, this term is merelyof conceptual importance because it is currently impossibleto measure reactive surface area. It is usually assumed thatthe reactive surface area value is a fraction of the total surfacearea. However, Lu¨ttge98 has raised the question whetherreactive and total surface area can even be compared directlysuggesting that the correct description of reactive surfacearea would require an energy term in addition to thegeometric term expressed in area units. In current rateequations, this reactivity term is usually “buried” in the rateconstant and may account for the wide range of valuesreported for the same mineral at similar conditions.103 In thiscontext, the development toward an often assumed quasi-steady-state surface may be questionable. These questionsshould motivate a reappraisal of our current approach thatwe take to quantify surface area.

3.2.5. Dependence on Distance from Equilibrium andScaling

After we have correctly quantified and applied the surfacearea to the normalization of reaction rate, any quantitativetreatment of processes involving fluid flow and mineralreactions needs to extract a crystal dissolution or growth ratelaw or both from these experimental data. This approach mustinclude the possibility of approaching mineral-fluid equi-librium. While many computer models routinely input suchrate laws,104 the experimental studies of mineral dissolutionreactions have not provided definitive results in manycases,105 except for conditions that are quite far fromequilibrium. Unfortunately, in natural systems, such far-from-equilibrium conditions are not very common. For mineraldissolution, this situation would be represented by the so-called “dissolution plateau”, where the dissolution rates ofminerals become usually independent of the difference inGibbs free energy (∆G) (for further information see, La-saga47). However, extrapolation of laboratory data to naturalconditions necessitates a description of the kinetic changesexpected as conditions approach near-equilibrium conditions.Recently, Beig and Lu¨ttge,106 Dove,107 Hellmann,108 andLuttge109 have presented new experimental results andtheoretical discussion seeking to address this topic. However,the application of laboratory data to natural systems remainsoften difficult or even unsuccessful.110-112

With the∆G dependence of the dissolution reaction rate,the surface area quantification, and our common struggle ofhow to link experimental observations at atomic or near-atomic scales to bulk dissolution rates, we have identifiedthe major reasons for potential failure in transferring labora-tory data to natural systems. The latter problem results fromthe fact that results of interface processes at the molecularscale can be observed at a variety of different time and lengthscales. And independently of the scale at which the ratesare measured, they need to be extrapolated, compared, orapplied to a largely different scale. Therefore, a major goalmust be to generate an understanding of reaction kineticsthat it is fundamental enough to provide the theoreticalconcept that allows us to link the processes and their ratesover the entire range of scales, both in time and length.Ultimately, this will be the prerequisite for a generalapplicability of laboratory data to natural systems.

3.3. Precipitation of Carbonate Minerals fromSeawater and Related Solutions

3.3.1. General Remarks

There is a substantial theoretical and observational un-derpinning to the general understanding of precipitationreactions available from the crystal growth literature. Thisfocus on precipitation and growth derives from the obviouspractical value of controlling reaction yields in industrialcrystallizers and chemical process reactors. However, ap-plication of this understanding to precipitation reactionsinvolving carbonate minerals in seawater has been limited.The reasons for this should be at least partly evident fromthe discussions in the previous section on solubility: thebehavior and reactivity of aragonite and magnesian calcitephases in seawater are strongly influenced by as yetincompletely understood surface reaction mechanisms andpotential formation of “solubility-controlling” surface phasesthat differ in reactivity from the bulk mineral, making evensimple seawater solubility determinations problematic. Be-

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 353

Page 13: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

cause of the nature of the relationship between free energyand surface-controlled dissolution and precipitation ratesintroduced in the preceding subsection, progress has ingeneral been greater in understanding carbonate mineraldissolution versus precipitation, in simple solutions versuscomplex brines, in part reflecting the significant laboratorychallenges in precipitation work due to the difficulty incontrolling reaction path.

Second, the bulk of carbonate precipitation in seawater isskeletal, and organisms may orchestrate processes of pre-cipitation and crystal growth (biomineralization) throughsophisticated regulation of an internal chemistry that maydepart significantly from the “constant ionic medium” ofseawater. Marine calcifiers may thus be able to accomplishskeletal growth in seawater solutions at rates that departsignificantly from those predicted by the “baseline” processunder strict abiotic control.

Third, both precipitation and dissolution reactions inseawater are also strongly influenced by inhibition. Althoughquantifying this inhibition in terms of net reaction rate maybe straightforward, a fundamental understanding of thisprocess in seawater ultimately requires knowledge of theactual inhibition mechanism. This knowledge in turn requiresinsight into the role of free energy (i.e., lattice ion activities)as a driving force versus that of specific dissolved or surface“impurity” species. Because growth surfaces inherit their ownsubstrates, precipitation reactions reflect dynamic controlsimposed by feedback interactions between mineral surfaceconfiguration and seawater solutes. It is only relativelyrecently that sufficiently sophisticated thermodynamic modelshave been available that simultaneously accommodate thestrong seawater electrolytes, the potential inhibitors that maybe present over a large concentration range, and the additionalcomplexities imposed by the carbonic acid system itself.Below, we adopt a “divide and conquer” strategy in an effortto parse what is understood about the individual roles of theprincipal physical parameters and chemical components thatexert fundamental control over the rate and fate of carbonatemineral precipitation.

3.3.2. Saturation State Controls and Their Relationship toInhibition

As described in the previous section, seawater saturationstate,Ω (expressed as IAP/K or ICP/K′, where ICP) ionconcentration product), occupies a central role in the rateequations for carbonate mineral precipitation. We should becareful to distinguish between controls on homogeneous andheterogeneous nucleation and precipitation. Homogeneousnucleation from solution occurs in the absence of pre-existingsurfaces and thus must involve initial formation of precursorphases whose assembly and growth will progress toward theproperties of the bulk crystal. The supersaturation requiredfor sustaining this process is described by the Gibbs-Thompson (also Ostwald-Freundlich) equation (discussedbelow), representing a balance between the decrease in freeenergy incurred by dint of the phase transition itself versusthe penalty of work (surface tension) expended in theconstruction of new surface area. With increasing supersatu-ration, the minimum (critical) radius of nuclei that cansurvive subsequent disassembly decreases. We will discussthe possible controls on nucleation rates below. In contrast,heterogeneous precipitation, such as an overgrowth, involvesaddition of material to a preexisting (large) surface.

The action of inhibitors greatly complicates the relationshipbetween free energy and the observed rate; rate equations

may attempt to isolate this effect by casting inhibitorconcentrations outside the∆Gr term (e.g., eq 3.8). However,we should note that there is a gap between what can beincontrovertibly demonstrated in terms of the mechanisticrole of impurities and what model assumptions are. In theclassic crystal growth model of Cabrera and Vermilyea79

described previously, adsorption of impurity ions at stepedges or terraces blocks proper step advance until sufficientsupersaturation is available and step motion can resume: steppinning is what is actually observed. Adsorption of impurityions may decrease the difference in chemical potentialbetween the solid and the solution, thus depressing thedriving force for reaction. Depending on saturation state,growth (and dissolution) processes may not be activeeverywhere on the surface, but limited to high-energy sites;competitive impurity attachment at these sites can effectivelyshut down the net reaction even at low relative coverage. Ifimpurities are incorporated with the growing lattice, thisreduction can be accomplished through the generation ofdefects (e.g., misfits or dislocations) that destabilize thesurface through an increase in the local strain field. Alter-natively, dissolved impurities may simply decrease the IAPthrough homogeneous formation of a ligand involving agrowth ion (e.g., MgCO30 formation).

These potential complexities between inhibition andsaturation state are manifest in both precipitation anddissolution reactions. If introduction of the inhibitor doesnot alter the actual reaction mechanism (reflected in areaction order with respect to ion activity product or a givencomponent thereof that is independent of inhibitor concentra-tion), the observed rate will still obviously suffer. In thecanonical expression of the seawater precipitation rate, thatis

introduction of inhibitors (magnesium, sulfate) may changethe rate constant (k) but not the reaction order (n). Alterna-tively, variation in inhibitor concentration may give rise tolittle change ink but produce significant changes in reactionorder (e.g., phosphate). Our point here is that the overallcompositional sensitivity of reaction rate parameters is afunction of the specific inhibitor. We discuss these effectsseparately in the sections below.

3.3.3. The Effect of Salinity and Ionic Strength

The effect of salinity and ionic strength on calciteprecipitation rate has been extensively investigated in avariety of solutions.113,114Relatively recent results are avail-able from Mucci and co-workers. In synthetic seawater,Zhong and Mucci115 found that calcite precipitation rateswere essentially invariant as a function of salinity (S) 5-44,Ω > 2.6, pCO2 ) 10-2.5), permitting a single fit to eq 3.12.As an aside, Mucci8 also investigated the effect of salinityon the stoichiometric solubility products, finding that thesolubility products of both aragonite and calcite tend toincrease with increasing salinity in a similar fashion. Zhongand Mucci115 also found that aragonite precipitation ratestended to decrease (2.5-5×) with increasing salinity,although this decrease was only observed forSg 35. Furtherevaluation of the ionic strength sensitivity of calcite pre-cipitation rate was extended in subsequent work by Zuddasand Mucci.116 In this work, they also attempt to formulate arate law on the basis of carbonate ion and describe calcite’s

r ) k(Ω - 1)n (3.12)

354 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 14: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

net precipitation rate as a balance between forward (f) andbackward (b) reactions,

Zuddas and Mucci reorganized this expression to yield afunction of carbonate ion concentration only:

whereKf contains both activity coefficients for calcium andcarbonate ion and is thus a function of ionic strength. Thisseparation of terms showed that as ionic strength is variedover a large range (0.1 mols/kge Iz e 0.93 mols/kg), thereaction order with respect to carbonate ion,n2, and theforward reaction rate constant,Kf, both increase. Theseexperiments also varyIz in NaCl-CaCl2 solutions and thusshow the effect of ionic strength independent of the inhibitingeffect of magnesium and sulfate (see below). Zuddas andMucci116 argued that the positive effect of ionic strength onprecipitation rate couldnot be understood as a simpleconsequence of a decrease in the energetic costs of dehydra-tion associated with decreasing water activity; the decreaseof aw is too small as ionic strength increases.

3.3.4. The Role of TemperatureVariation in temperature affects precipitation reactions in

two fundamental ways. First, temperature is assumed toinfluence fundamentally the reaction rate constant throughthe classical Arrhenius equation,

Here A is a prefactor term (assumed to be independent oftemperature), andεA is the activation energy. At the risk ofsome repetition, we emphasize that precipitation rates asmeasured in seawater reflect overall, and not elementary,reactions. Although some observational constraints exist, thereaction mechanisms by which, for example, a magnesiancalcite, aragonite, or protodolomite forms on a molecularbasis are largely unknown. For magnesian calcite, variationsin temperature could potentially give rise to variations inreaction mechanism, leading to respective differences intemperature dependency (εA) as well.

Second, mineral solubility and the equilibrium distributionof dissolved species (ion activity product) are both temper-ature-dependent, and thus temperature variations even atfixed composition give rise to variations in the driving forcefor the reaction as well. The solubility of all pure carbonatesdecreases with increasing temperature. In the case of aprecipitating magnesian calcite, temperature also affects theextent to which magnesium dissolves in solid solution (seepreceding section on the relationship of magnesium contentto calcite solubility). Early experiments all showed positiveincreases in magnesium content of calcite with respect totemperature.117-120 As mentioned previously, the compositionof magnesian calcite overgrowths formed in 25°C seawaterdepended solely on Mg/Ca solution ratio and was indepen-dent of rate or seawater saturation state.24,26The explicit roleof temperature in controlling overgrowth composition andrate of magnesian calcite precipitation was subsequentlyexplored in Mucci.121 When plotted versus temperature(Figure 6), these results confirm a positive temperaturedependence on the extent of magnesium uptake, with a

similar slope as that observed in previously cited work,although with significant variation in intercept. With theexception of the results of Fuchtbauer and Hardie,120 thesesynthetic compositions also lie outside the domain of thecommon magnesian calcite cements in modern oceans,leading Mucci121 to conclude that temperature variations areof insufficient strength to explain the natural distribution ofabiotic marine calcite cement composition. However, tem-perature may have a more important effect in determiningthe mineralogy of the phase that initially nucleates inseawater; this aspect is discussed in section 3.3.8.

3.3.5. The Roles of Magnesium and Sulfate

As discussed in the previous sections, the effects ofmagnesium, although well-documented, are complex andpoorly understood from a mechanistic standpoint. Wesummarize those effects here in terms of the relevance toprecipitation reactions:

1. Lattice incorporation of magnesium in bothnaturally occurring abiotic and biogenic calciteresults in a more soluble phase, provided thatsolubility is computed according to eq 2.2; however,this increase may be accompanied by little relativechange in the activity product of calcium andcarbonate ion. In seawater, magnesian calcitecomposition appears to be primarily controlled bythe Mg/Ca ratio of seawater and is otherwise largelyindependent of saturation state and precipitationrate. If variations in MgCO3 composition exist in agiven overgrowth, these differences may driveselective, time-dependentdissolutionof more solublecompositions during subsequent equilibration epi-sodes.45

2. At a given mole percent of MgCO3, abiotic(synthetic and marine cement) phases are less

r ) kf aCa2+n1 aCO3

2-n2 - kb (3.13)

log(r + kb) ) log Kf + n2 log [CO32-] (3.14)

k ) A exp(-εA/(RT)) (3.15)

Figure 6. Variation of MgCO3 content of magnesian calcite withtemperature, precipitated in seawater or seawater-like solutions,adapted from Mucci121 (see Mucci121 for references to original data).Solid squares are CaCl2 + MgCl2 + NaCl; solid circles are seawater+ Na2CO3 + Ca(HCO3)2; open circles are seawater+ Na2CO3;open triangles are CaCl2 + MgCl2 + Na2CO3; open squares areseawater+ Na2CO3 + aragonite seeds; solid triangles are seawater+ calcite seeds. Reprinted from ref 121, Copyright 1987, withpermission from Elsevier.

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 355

Page 15: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

soluble than biogenic marine calcites. The work ofBischoff and co-workers21,22,122and Gaffey123 dem-onstrates that regardless of magnesium content,biogenic calcite contains variable water and exhibitsmore carbonate ion and cation disorder comparedwith synthetic phases. In an extensive review,Morse and Mackenzie1 concluded that the structuraland compositional heterogeneity of biogenic mag-nesian calcites, together with additional complica-tions such as Mg(OH)2 inclusion,124 in generalwarranted their exclusion from conventional deter-minations of thermochemical stability.

Although our understanding of the mechanism of mag-nesium incorporation is sufficiently limited as to precludedefinitive statements, there is some evidence that magnesiummay play a role in terms of introduction of water into thecalcite structure. For example, Lippmann125 suggested thatwater molecules may be incorporated in magnesian calcitedue to the large amount of energy required to completelydehydrate the Mg2+ ion during precipitation. The incorpora-tion of water and its possible correlation with magnesiumcontent is also consistent with spectroscopic data,16,123leadingto the hypothesis of a “hydrated magnesian calcite”.1 Excesswater may also exist in biogenic phases poor in MgCO3,suggesting that calcium ion may also be partially hydrated.This mechanism would also be temperature-sensitive, be-cause complete dehydration and incorporation of anhydrouscalcium and magnesium ions into the solid would becomeincreasingly difficult with decreasing temperature.121,126,127

In contrast to its affinity for calcite, magnesium has asignificantly lower affinity for the aragonite surface. Mucciand Morse128 were able to use profiling Auger spectroscopyto demonstrate that in supersaturated seawater, magnesiumshows more than an order of magnitude less adsorption onaragonite surfaces compared with that on calcite. In addition,although increasing saturation state (Ω) leads to a slightincrease in the adsorbed Mg2+/Ca2+ ratio of the calcitesurface, increasingΩ has a strong negative effect on thearagonite surface ratio.

Sulfate is recognized as a significant coprecipitating anionin marine biogenic calcites (∼1 mol % in magnesiancalcites), and its effect on calcite precipitation has beeninvestigated in solutions similar to seawater.129,130 In Mg-free solutions, Busenberg and Plummer129 found sulfateincorporation by calcite tended to increase with the rate ofprecipitation; increasing sulfate activity in turn inhibited theprecipitation rate, with increasing inhibition as a functionof saturation state. By comparison, relatively little sulfate istaken up in aragonite, and natural marine aragonites typicallycontain less than 6000 ppm SO4

2-.129

Last, we do not discuss dolomite reactions in this paperbecause of their limited importance in modern sediments incontact with normal seawater. However, experimental workby Brady et al.130 may have general relevance to reactionsinvolving carbonate surfaces in the context of interactionsbetween magnesium and sulfate. This work, involving cationsorption on dolomite, showed that the uptake of bothmagnesium and calcium to the mineral surface is potentiallyenhanced by the addition of dissolved sulfate. At themoderate pH values of seawater, this behavior was thoughtto reflect co-adsorption of metal sulfate complexes; in thecase of magnesium, the uptake of sulfate could displacewaters of hydration of adsorbed magnesium ion, providing

a pathway for carbonation and subsequent lattice incorpora-tion.

3.3.6. The Role of Other Impurities: Manganese,Strontium, Phosphate, and Organic Matter

There has been considerable effort devoted to understand-ing the role of manganese in calcite precipitation reactions.In general, this work has shown that Mn2+ is adsorbedstrongly on the calcite surface, even at concentrations wellbelow rhodochrosite (MnCO3) solubility. Mn2+ is partitionedinto solid calcite (DMn > 1), although it has been demon-strated that this partitioning is sensitive to precipitation rate,with increasing rate producing a reduced partition coef-ficient.131,132

Early work in dilute solutions demonstrated that atsufficient surface concentrations, initial chemisorption ofMn2+ is followed by nucleation of MnCO3.133 Reactionsinvolving Mn2+ and calcite were explored in seawater byFranklin and Morse,134 with the finding that the interactionof Mn2+ with calcite is strongly influenced by Mg2+

availability. In magnesium-free seawater, the pattern issimilar to that observed in dilute solutions, with an initiallyrapid period of Mn2+ uptake, followed by MnCO3 nucleationand subsequent growth that is first order with respect todissolved Mn2+ concentration. Franklin and Morse134 ob-served that the introduction of magnesium at seawaterconcentrations apparently interfered significantly with Mn-CO3 nucleation, possibly through site competition, thusslowing the growth rate of manganese carbonate. In addition,attempts to measure rhodochrosite solubility in seawater fromsupersaturated conditions resulted in formation of a mixedsolid. Subsequent work by Mucci135 confirmed the inversedependence on precipitation rate observed in previous work,with manganese carbonate forming as a (Ca,Mn)CO3 solidsolution containing up to 40 mol % MnCO3. However, thisdependence of solid-phase manganese composition on pre-cipitation rate precludes the condition of exchange equilib-rium suggested above in explanation of the relationshipbetween seawater and magnesian calcite composition. In fact,Mucci135 argued that the overall precipitation rate of the solidcould be successfully modeled as a cumulative function ofthe end member rates. In addition, although manganesecontent of the overgrowths decreased in response to increas-ing precipitation rate, the Mg/Ca ratio was constant (againconsistent with earlier results for magnesian calcite). Mucci135

suggested that composition of this mixed phase pseudokut-nahorite137 was consistent with the observed composition ofmanganoan carbonates derived from reducing pore waters.

The interaction of strontium with calcite and aragonite hasbeen the focus of numerous experimental studies. Lorens131

showed that at seawater salinities, the observed Sr partitioncoefficient in calcite is strongly dependent on precipitationrate, increasing by a factor of∼3 as rate increases by a factorof 103. Mucci and Morse24 also observed that Sr uptake inmagnesian calcite overgrowths was linearly related to theMgCO3 content, possibly reflecting changes in lattice dimen-sion. Morse and Bender6 summarized these data in thecontext of earlier experimental results, and showed thatSrCO3 content is clearly correlated with variation in ratedespite additional increases in the Sr content of magnesiancalcites. Strontium accommodation in aragonite is much moreextensive than in calcite, reflecting lattice differencesdescribed earlier. Plummer and Busenberg136 have investi-gated the strontianite (SrCO3)-aragonite solid solution in

356 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 16: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

extensive experiments and thermodynamic modeling. Theyconcluded that although marine aragonites (∼1 mol %SrCO3) have compositions close to the maximum stabilityobserved, excess Sr in these phases suggests that they are inneither stoichiometric saturation (see discussion above forthe magnesian calcites) nor equilibrium with respect to theseawater strontium concentration.

Orthophosphate adsorbs strongly to calcite and aragonitesurfaces in seawater, exhibits complex interactions with thecarbonate surface, and generally strongly inhibits carbonateprecipitation. In dilute solutions, Stumm and Leckie137,138

observed a period of rapid initial uptake and a resting period,followed by a period of renewed uptake, indicating chemi-sorption, recrystallization-nucleation, and growth of apatite.In seawater, the presence of magnesium has been shown toinhibit apatite formation.139 In seawater experiments, deKanel and Morse140 observed that the rate of phosphateuptake to the calcite and aragonite surfaces actually declinedwith time. Although the initial uptake of phosphate wasgreater for aragonite than calcite, the subsequent decline incalcite uptake rate was slower than that for aragonite. Thesetrends are inconsistent with a Langmuir adsorption isothermbut may be explained by either an exponential decrease inthe availability of reactant sites or a linear increase in theenergy of activation related to phosphate uptake. Mucci126

modeled the strong inhibition of phosphate as a shift inreaction order, fromn ) 2.8 in phosphate-free seawater toa constantn ) 3.4 in the presence of phosphate, and as avariation in the rate constant itself, thus amending thelogarithmic form of the canonical rate equation as

where a + log a + b log [PO43-] represents a linear

expansion of the rate constant in phosphate concentration.Last, the interaction of organic matter with carbonate

mineral surfaces is significant and diverse. For example, inearly experimental work (discussed later in section 3.6.3),Suess141 showed that calcite exhibits a strong but selectiveaffinity for dissolved organic matter and suggested thatadsorption to the point of formation of monomolecularsurface layers could result in chemical isolation of carbonateparticles. Berner et al.142 have demonstrated that introductionof carboxylic, humic, and fulvic acids will inhibit nucleationof aragonite in supersaturated seawater. Zullig and Morse143

examined the adsorption of fatty acid homologs (butyrate,octanoate, laurate, myristate, palmitate, and stearate) oncalcite, dolomite, aragonite, and magnesite, finding thatsurface affinity increased with alkyl chain length, such thatlittle adsorption was observed for chains shorter than C14

and essentially quantitative and irreversible adsorption forC18 acids. This pattern is essentially the inverse of fatty acidsolubility. This behavior is consistent with the interactionof these hydrophobic acids with the carbonate surface viathe carbon chain itself, as opposed to the interaction betweenthe exposed surface calciums and carboxyl groups.

3.3.7. Nucleation of Calcium CarbonateThis description is similar to that available in Nielsen;54

treatments found in related texts will also be similar. In orderto describe the dependence of nucleation on saturation state,we begin with the critical assembly of molecular units to asmall cluster. If we assume that the changeper moleculeinchemical potential from the dispersed to assembled state is

µ2 - µ1, then the total free energy change forn moleculesof material is

or per molecule

where the left-hand quantity (-∆G/n) must be>0 for aspontaneous assembly to occur. If we now approximate thevolume of this cluster ofn units as a sphere having volumeV, the surface areaA is proportional ton2/3. Thus the freeenergy change for the homogeneous formation of a crystalassembly involvingn units from solution is

or letting ψ ) Ασ/n2/3 express the proportionality betweensurface work (σ) and the number of molecules in the crystal,we now have as a function of crystal size (or number ofassembled molecules)

Equation 3.20 is plotted for a range of∆µ values in Figure7 and represents the free energy change as a balance betweenthe benefit of assembly (the phase change itself) and the costof increasing surface area, scaled by surface tension, as thenumber of assembled units,n, increases.∆G is positive for∆µ < 0 (undersaturation) and becomes infinitely so withincreasingn. However, under supersaturated conditions (∆µ> 0), although∆G will be positive in the region of smalln,becausen grows faster thann2/3 as n increases,∆G willeventually reach a maximum value and become increasinglynegative thereafter. This maximum is an equilibrium point,although an unstable one, and may be computed from thederivative

log r ) a + b log [PO43-] + n log(Ω - 1) (3.16)

Figure 7. Variation of free energy (∆G) versus crystallite size(radius), computed as a function of increasing oversaturation (∆µ)of the solution, using eqs 3.19 and 3.20. Open circles representequilibrium condition, light gray circles representΩ ) 2 (arrowindicates position of critical radius), dark gray circles representΩ) 5, and black filled circles representΩ ) 50.

∆G ) n(µ2 - µ1) (3.17)

-∆G/n ) µ1 - µ2 ) -∆µ (3.18)

∆G ) -n∆µ + σA (3.19)

∆G ) -n∆µ + n2/3ψ (3.20)

n* ) (2ψ/(3∆µ))3 (3.21)

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 357

Page 17: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

This numbern* thus defines the critical radius: crystals withradii smaller than this size (fewer molecular units) will tendto disintegrate, and ones that become larger thann* will tendto grow (Ostwald ripening). As can be seen in Figure 7, thecritical size is also a function of supersaturation: the higher∆µ, the smaller a given crystal assembly need be to surviveand grow.

Early experimental studies of the conditions under whichcalcium carbonate can undergo pseudohomogeneous nucle-ation from seawater have been conducted that indicatedcalcium carbonate can nucleate during seawater evaporation(for summaries see Braitsch144 and Sonenfeld145) and thatincreasing salinity should promote calcite nucleation. Studiesof calcium carbonate nucleation from seawater with salinitiestypical (S ≈ 35) of most oceanic waters demonstrated thatnucleation was promoted by increasing salinity and decreas-ing temperature.146 Results obtained by Pytkowicz,147 asrecalculated by He and Morse,148 showed that a supersatu-ration of about 20-30 times with respect to calcite wasnecessary for nucleation of calcium carbonate within a fewhours to a day. Subsequently, Morse and He149 conductedexperiments to investigate the influences of pCO2, temper-ature, and salinity on calcium carbonate nucleation fromseawater. They found only modest influences of temperatureand salinity on the degree of supersaturation (about 20-25times with respect to calcite) necessary for calcium carbonatenucleation to occur from seawater as shown in Figure 8. Thisis about 3-4 times the typical supersaturation of surfaceseawater.

The nucleation of calcium carbonate from seawater hasalso been of considerable interest from the standpoint ofwhich mineral is precipitated. At 25°C and close to normalsalinity (35), aragonite is nucleated and precipitated fromseawater even though it is unstable relative to calcite ordolomite. However, observations of ancient sediments in-dicate that at times in the past, when the composition ofseawater may have differed from that of modern seawater,particularly with respect to lower Mg/Ca ratios, calcite wasprobably favored over aragonite.150-152 Also, natural observa-tions indicate that in cold waters calcite is more likely toform than aragonite. The literature on this topic has beenreviewed by Morse et al.153 who also conducted experimentalstudies of factors controlling which calcium carbonate phasenucleates and grows from seawater under differing composi-tion and temperature conditions. Their major results aresummarized in Figure 9. Calcite forms over a wide temper-ature range (0-35 °C) in Mg-free seawater, but only belowabout 8°C in seawater of normal (∼5:1) Mg/Ca ratio. From

these results, it is evident that both the Mg/Ca ratio andtemperature control the calcium carbonate mineral that formsfrom seawater, effectively dividing the resulting “phasespace” between calcite and aragonite.

Another interesting aspect of this work is that Mg/Ca ratioand temperature control not only which calcium carbonatepolymorph precipitates but also the morphology of theprecipitate itself (Figure 10). In some experiments calcitewas observed to nucleate initially, but as precipitationproceeded the solution Mg/Ca ratio increased until aragoniteepitaxially nucleated on the calcite, followed by growth ofaragonite needles (Figure 10C).

Figure 8. The supersaturation with respect to calcite (Ωcalcite)necessary for pseudo-homogeneous nucleation of calcium carbonatefrom seawater at different salinities. Red line and symbols are at25 °C, and blue line symbols are at 40°C. Modified from ref 149,Copyright 1993, with permission from Elsevier.

Figure 9. Influences of seawater Mg/Ca ratio and temperature onthe nucleation of calcium carbonate from seawater. Blue circlesare aragonite; yellow circles are calcite; green circles are initiallycalcite with aragonite overgrowths. Modified from ref 153 withpermission. Copright 1997 Geologic Society of America.

Figure 10. Scanning electron photomicrographs of calciumcarbonate precipitated from seawater of various Mg/Ca ratios andtemperatures: (A) aragonite with needle morphology from warmnormal sweater; (B) aragonite with “broccoli” morphology fromcold normal seawater; (C) aragonite needles growing on initiallynucleated rhombohedral calcite for low Mg/Ca cold seawater; (D)radially fibrous aragonite sphere from 15°C normal seawater.Panels A, B, and C reproduced from ref 153 with permission.Copyright 1997 Geologic Society of America. Panel D photo takenby J. Morse.

358 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 18: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

3.4. Dissolution Reactions

3.4.1. Calcite Dissolution KineticsThere are significant differences between dissolution and

precipitation reactions, and although approaches that cast thetwo regimes in the same theoretical light may yield in-sight,154,155in general we would argue that dissolution cannotbe understood simply as crystal growth operating with anegative sign. This limitation is important in the context ofseawater reactions, where our lack of quantitative under-standing of reaction paths and the role of “precursor” surfacephases is particularly acute.

Although there is a large experimental literature oncarbonate mineral dissolution, much of this work has beendone at extreme undersaturations and low pH regimes thatare in the province of diffusion-controlled kinetics and havelimited relevance to reactions in seawater. Under theseconditions, dissolution rate may reflect primary control bythe hydrodynamic conditions that maintain the thickness ofthe stagnant boundary layer. For example, under acidicconditions (pH< 4), reaction rates can be described simplyby (compare also eq 3.8),

wheren is ∼1 and the rate constant (25°C) varies between50 × 10-3 cm s-1 147 and 3× 10-3 cm s-1.156 The range ofthese values reflects the uncertainty and sensitivity toestimates of the thickness of the boundary layer. The progressto equilibrium from a purely transport-controlled regime ischaracterized by a transitional region. This transitional region,in which reaction rates may be intermediate between strongdependence on hydrodynamic conditions (rotating diskvelocity) and complete surface control, was the focus ofmuch work by Sjo¨berg and Rickard156-158 and gave rise tothe notion of “mixed kinetics”. This transitional behaviorwas described by expressing the observed rate constant as anonlinear function of a transport-controlled (kT) and chemi-cally controlled (kC) rate constant. Thus ifkT . kC, the rateis chemically controlled; conversely forkC . kT, the rate istransport controlled:

Once surface-controlled dissolution kinetics are established,reaction mechanisms vary through this near-equilibriumregion until equilibrium is attained. These relationships aresummarized in Figure 11. We also refer the reader to a recentand relatively exhaustive review of the carbonate mineraldissolution literature.50

Although there is a general consensus regarding the criticalimportance of understanding surface-controlled reactions inthe limited range over which saturation states vary in normalseawater, much fundamental knowledge is lacking. Themajority of studies of dissolution in seawater consist ofapplication of painstaking analysis of solution chemistry asa means of deriving insight into activities on the surface itself.The results of this work have provided a large dataset thatdescribes the overall relationship between bulk solutioncomposition and reaction rates. However, mechanistic in-terpretation of surface reactions is limited by a lack of directsurface observations and measurements. In addition, thereis the more fundamental problem of how to integrate theenormous detail derived from current AFM studies of mineralsurfaces over significant length and time scales to providenet reaction rates relevant to mesoscale processes. Thus amajor outstanding challenge is the integration of these variousapproaches and observations to yield models of practicalutility.

The overall relationship of temperature to calcite dissolu-tion rate reflects the differences between transport versussurface-reaction control described above. Early work, sum-marized by Plummer et al.,159 shows substantial variation inactivation energies, although reliable values in dilute solu-tions are similar (e.g., 10.5 kcal mol-1 160). Experiments bySjoberg and Rickard158 far from equilibrium show that therole of temperature on calcite dissolution kinetics is complexand are summarized in Figure 11. Closer to equilibrium,Gutjahr et al.161 confirmed surface control of calcite dissolu-tion with an activation energy of 8.3 kcal mol-1. Thetemperature dependency of dissolution inhibitors is poorlyknown close to equilibrium.

The role of inhibitors in dissolution reactions has generallybeen best described for conditions far from equilibrium indilute solutions. These results are reviewed in more detailin Morse and Arvidson50 and only summarized briefly here.

1. Although substantial experimental observationsexist for inhibition of dissolution reactions, thevariation in terms of saturation state and major ioncomposition makes it difficult to draw overallconclusions regarding the role of specific inhibitors.This is important because the degree of inhibitionis often quite sensitive to the distance from equi-librium.

2. As described previously in the section onprecipitation, inhibition is often described accordingto competing models: (a) a reversible adsorption(Langmuir-Volmer) model, which assumes that aLangmuir adsorption isotherm controls uptake atspecific surface sites, or (b) an irreversible adsorp-tion (Cabrera and Vermilyea) model in which thedistribution of impurities on terraces controls theadvance of steps. Control by either of these twomechanisms is often simply assumed and notclearly demonstrated.

3. In addition to impurities, lattice components(calcium and carbonate ion) may also influencereaction rate beyond their simple appearance insaturation state terms.160 In addition, the role ofinhibitors in seawater is also influenced by calciumand magnesium concentrations. Other than sulfateand phosphate, Sjo¨berg160 observed that other majorand minor ions of seawater have little influence oncalcite dissolution rate.

Figure 11. Summary diagram of relationship of transport versussurface-controlled dissolution based on the work of Sjo¨berg andRickard.162 Modified from ref 50, Copyright 2002 with permissionfrom Elsevier.

r ) kaH+n (3.22)

k ) kCkT/(kC + kT) (3.23)

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 359

Page 19: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

4. Trace metals (e.g., Ni, Cu, Mn, or Co) exertstrong control on calcite dissolution even at mi-cromolar concentrations, possibly by selectiveblocking of active sites.80 The effectiveness of thesemetals may also be dependent on saturation state,but they have not been extensively studied inseawater where they may be less influential due tosite competition.

Early work involving in situ Pacific Ocean experi-ments163,164 related dissolution rates to seawater saturationstate. This critical work showed that the seawater dissolutionrate of calcite is not simply proportional to undersaturation,obviating the use of first-order kinetics (see above). Similarresults were also obtained for the Atlantic Ocean.165 Morseand Berner49,166,167performed extensive experiments over abroad range of saturation states, using∆pH (the differencebetween the equilibrium and experimental pH) to describesaturation state. The primary difficulty in the study ofseawater dissolution kinetics of calcite is the small variationin seawater saturation state (Ω typically >0.7 with respectto calcite). This thus limits the range of relevant undersatu-ration (∆pH e ∼0.2). Using the canonical rate law writtenfor dissolution, the observed reaction order (n) for phosphate-free seawater is∼3 for inorganic calcite powders166 and 4.5for biogenic calcite tests.49,168This work was critical in termsof establishing the increase in dissolution rate associated withan undersaturation close to that observed for the foraminiferallysocline in deep sea sediments.166 In addition, Morse49

identified phosphate as the most important inhibitor in termsof calcite dissolution in seawater. Although the presence ofseawater phosphate exerts a minor influence on the rateconstantk, reaction ordern is increased to∼16 at 10µMphosphate. Berner and Morse167 argued that selective bindingof phosphate at high-energy sites progressively raised the“critical undersaturation” necessary to maintain step retreat.

As with precipitation reactions, understanding of dissolu-tion is critically hampered by fundamental gaps in theunderstanding of how microbial processes, both in the watercolumn and in sediments, exert controls over dissolutionreactions. In addition, application of rate laws such as thosedefined previously to actual dissolution in marine environ-ments is handicapped by an inadequate understanding of howto extrapolate these terms in time and space. This problemis covered in more detail in section 7.

3.4.2. Aragonite Dissolution

Compared with calcite, the attention directed at aragonitedissolution in seawater has been relatively modest. Morseet al.169 showed that synthetic aragonite and pteropod testsdissolved under conditions of near-equilibrium toΩ ) 0.44with an overall reaction ordern ) 2.93, similar to thatobserved for calcite. With increasing undersaturation,nincreased to 7.27. The behavior of phosphate is complex anddiffers substantially from that observed for calcite, in thatintroduction of phosphate gives rise to an initial increase indissolution rate. However, this effect is dependent on bothsaturation state and phosphate concentration169 and is alsotime-dependent.

In subsequent work, Busenberg and Plummer170 showedoverall similarities between reaction kinetics and rate con-stants determined for aragonite and calcite; these rates weretreated according to a substantially different set of rates thanpresented above and attempt to assign mechanistic roles tocarbonate surface species. Gutjahr et al.161 derived a reaction

order for aragonite of∼3, in good agreement with the earlierwork of Morse et al.169 in seawater.

3.4.3. Magnesian Calcite Dissolution Kinetics

We have previously discussed the problems related to thedetermination of the solubility of magnesian calcite. In earlywork, Land171 observed that biogenic magnesian calcitedissolved in pure water incongruently, which he arguedreflected a faster MgCO3 versus CaCO3 dissolution rate.Subsequent work by Plummer and Mackenzie17 essentiallyused the dissolution of magnesian calcites to derive equi-librium constants from reciprocal time. Plummer and Mack-enzie17 concluded that although the initial dissolution wascongruent, the dissolving phase differed in composition fromthe bulk solid. They argued that this behavior was controlledby dissolution of a range of magnesian calcite mineralogies,with variations in solubility producing variations in rate.Wollast and Reinhard-Derie172 confirmed these results insimilar free drift experiments, including reprecipitation of acalcite poorer in Mg than the initial reactant. Walter andMorse18 have also shown that the reaction ordern formagnesian calcite dissolution computed from rate versus1 - ΩMgCO3 (i.e., stoichiometric saturation) are similar topure calcite. Bertram et al.31 showed that the solubility ofsynthetic magnesian calcites decreases with temperature ina manner similar to pure calcite.

Walter and Morse173 measured dissolution rates of care-fully prepared biogenic magnesian calcites, computing rateconstants and reaction orders from stoichiometric activityproducts and stoichiometric constants.18 The importantobservation made in these experiments was that microstruc-tural differences exert the principal control on their dissolu-tion reactions, rendering determination of the role ofsolubility problematic or irrelevant. Rates were reported ona per unit mass (versus per unit area) basis because of largedifferences in reactive versus BET surface area. Walter andMorse18,173observed that the critical differences in rates werecontrolled not by solid-phase composition but by relation-ships between total surface area, grain size, and microarchi-tecture. Rates increased with BET surface area but not inany simple linear way. They found that complex microstruc-tures of biogenic Mg calcites may have apparent “reactive”surface areas ofe1% of total area. In these cases, BET areameasurements are an unreliable measure of reactive area.

In common with their behavior in supersaturated solutions,magnesian calcites exhibit unique properties that complicatedetermination of their reaction kinetics. As stated previously,an adequate representation of their kinetics hinges on accuraterepresentation of the distance from equilibrium. The use ofstoichiometric saturation in this regard may permit compari-son of phases of different composition, but it trades this utilityfor ambiguous theoretical constraints. In the case of biogenicmagnesian calcites, solution to this “solubility problem” mayrequire approaches entirely divorced from those employedin the description of synthetic or cement phases, in whichmagnesium content is only a minor factor. Despite a wealthof observations, conclusions regarding the actual dissolutionmechanism of magnesian calcites and the action of specificinhibitors remain largely speculative regardless of origin.Last, these properties (MgCO3 content, heterogeneity, grainsize, microstructure, defect density, etc.) need to be under-stood in terms of their implications for reaction path duringstabilization in natural environments.33

360 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 20: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

3.5. Biominerals and Nanophase Carbonates

3.5.1. General Considerations

Life imprints its signature upon Earth’s solid surface andthe oceans by causing and modifying important fluid-solidprocesses. The rapidly growing literature records the increas-ing number of attempts to understand the nature and role ofthis effect at scales ranging from global to microscopic.174-177

At the microscopic level, the interactions between microor-ganisms and minerals have repercussions that have a varietyof implications for the biogeochemical cycling of chemicalelements, the formation of minerals and the contrary, thatis, the dissolution and weathering of minerals and rocks andtherefore the formation of soils. Below, we will discussbriefly biomineralization, biogeochemistry and geobiologyof carbonates, and nanophase carbonatesonly with theirrelevance to seawater and oceanography.

3.5.2. Biomineralization

The term biomineralization, in contrast toabiotic mineralformation, describes the production of mainly crystallinematter by organisms, that is, crystal nucleation and growthgenerated by organisms or even inside an organism. Theproduct of this process is a biomineral. While in the pastEarth scientists were mainly focused on abiotic inorganicmaterial, that is, minerals per definition, this focus haschanged in the more recent past. Comprehensive reviews ofthis new and rapidly growing field of biomineralization havebeen provided by Lowenstam and Weiner178 and Dove etal.179 The interested reader will find a wealth of referencesto key research articles as well as to more specialized anddetailed reviews in these volumes. Dove et al. have as-sembled an interesting review volume that contains a dozenarticles. Together, this assembly provides a glimpse on thecutting edge of this field and an overview and great start toexplore this exciting subject in more detail.

In our present article, however, the focus is almostexclusively on abiotic carbonate minerals and their behaviorin oceanographically relevant settings. Therefore, we restrictourselves here to a few short remarks on biominerals andbiomineralization. Today, this field has developed so suc-cessfully and covers such a broad range of interdisciplinaryresearch that a satisfying coverage would require a fullreview article on this subject. Our decision is also supportedby the fact that biominerals are formed by organism-inducedprocesses that are very different from abiotic processesgoverned by abiotic nucleation, growth, and dissolutionkinetics. Due to the distinctly different way of formation thatoften involves organic molecules and matter, biomineralscommonly behave very different compared with their abi-otically formed counterparts. Thus certain aspects of ourdiscussion of abiotic carbonates may not apply to biomin-erals.

3.5.3. Biogeochemistry, Geomicrobiology

While we decided not to focus in detail on biomineral-ization (see above), it makes more sense to discuss theinfluence of microorganisms, such asShewanellaspp. andGeobacterspp., on the abiotic reaction kinetics in terms ofbiogeochemistry and geomicrobiology. Here, we definebiogeochemistry as geochemistry altered or influenced bybiological factors and accordingly geobiology as biologyinfluenced or altered by geology or geochemistry.

Bacteria can help formation of precipitation of carbonateminerals,180-182 but they can also do the opposite, trigger orenhance dissolution or corrosion. We will show below thatan intermediate action is also possible, that is, the inhibitionof dissolution or corrosion kinetics. For example, it has beenobserved in a relatively large number of studies183,184 thatbacteria will attach to metal, metal oxide, mineral, and othersolid surfaces. In many cases, this process will develop intothe formation of a biofilm (e.g., Escher and Characklis185),that is, a mixture of microorganisms (bacteria) and ex-trapolysaccharides (EPS) and other organic material that“grows” on the solid surface. It is easy to imagine that sucha complex biofilm will eventually alter abiotic reactionkinetics of dissolution or growth processes at the water-solid interface because it creates its own microenvironmentwith respect to the solution chemistry, pH, etc., or at least italters the pre-existing conditions significantly. Banfield andNealson183 have devoted an entire review volume titled:“Geomicrobiology: Interactions between Microbes andMinerals” to this interesting phenomenon. This work is anexcellent source for our understanding of microbe-mineralinteractions and contains a wealth of references.

While the complex interplays between bacteria, biofilm,and water-solid interface are still difficult to quantify andunderstand in detail, we have begun to make substantialprogress. Today, the forces and interactions between singlemicrobes and mineral surfaces can be measured,184,186 andwe have also started to elucidate the mechanisms that bacteriause to interact with the mineral surface. In this context,electron shuttles and nanowires are just a couple of recentobservations that are defining our progress. An improvedinsight and understanding of these mechanisms is importantfor several reasons. One example is the observation reportedrecently by Luttge and Conrad187 who demonstrated experi-mentally thatShewanella oneidensisMR-1 not only attachto calcite cleavage surfaces but also are capable of attachingto specific high-energy sites, that is, the outcrops of screwdislocations at the mineral surface (Figure 12). It is well-known from crystal growth and dissolution theory188,189thatthese high-energy sites are also key features in both crystaldissolution and crystal growth processes. They are both thesites where etch pit formation begins during dissolution andalso centers for the development of so-called growth spiralsthat form hillocks and later new crystal faces during thegrowth process. Lasaga and Lu¨ttge190,191have shown that etchpits are the main source for the formation of steps that willprogress192 and cause the surface-normal retreat of the entirecrystal surface, leading to overall crystal dissolution.

It has been suspected and demonstrated for some time nowthat microorganisms could cause or at least increase corrosionof metal and mineral surfaces.193,194The new observation thatmicroorganisms attach preferentially to high-energy sites isinteresting because this observation changes the assumptionof corrosion or dissolution enhancement in an important way.

Luttge and Conrad179 have observed that MR-1 organismsare capable of shutting down the abiotic dissolution processesof calcite even at significantly undersaturated conditions. Thereason is simply that the high-energy sites that become almostcompletely occupied by MR-1 do not open up into etch pits.Consequently, no steps are generated at the outskirts of thepits that would move across the crystal surface and woulddissolve the crystal layer by layer. However, individualorganisms would then begin to entrench themselves in thecalcite surface (Figure 13), but atomic force microscopy

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 361

Page 21: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

(AFM) and vertical scanning interferometry (VSI) imagesshow that the resulting pit is no longer crystallographicallycontrolled as in the case of etch pits. Instead, after MR-1has been removed from the surface, the pits show the oval,negative-body shape of theShewanellaorganisms (Figure14). The outer rim seems to be quite well controlled by MR-1; that is, it does not start to spread steps. The depth of thepits formed by MR-1 was usually no deeper than about athird of the organisms diameter, that is, about 140-200 nm.This result indicates that MR-1 can indeed control thedissolution process independent of the abiotic solutionchemistry in contact with the carbonate surface. Theseobservations are in close agreement with model resultspublished by Lu¨ttge et al.195 Davis and Lu¨ttge196,197 havereported that this behavior is not limited to calcite but wasobserved also for dolomite and magnesite, as well asrhodochrosite.198

Future work will need to be undertaken to substantiatethese still somewhat preliminary results. In particular, it willbe important to demonstrate that other microorganisms suchasGeobacteror at least other microbial strains ofShewanellacan operate in a similar manner. If more evidence will becollected, we may have identified an important phenomenonthat could explain unexpected or unexplained dissolution andgrowth rates in natural systems that are not in agreement

with expectations based solely on abiotically measured orestimated reaction rates.50,99

3.5.4. Nanophase Carbonates

An extensive literature search reveals that our recognitionand knowledge of abiotic nanoparticles in the carbonatesystem at least with relevance to seawater and oceanographyare mostly lacking. Most available articles are focused onbiologically formed nanoparticles that are not the subject ofthis review (see above). Banfield and Navrotsky199 havepublished an interesting and comprehensive review volumeon nanoparticles and the environment. This volume provides

Figure 12. (A) Reflected light image showing colonies and singleindividuals of Shewanella oneidensisMR-1 with superimposedinterference fringes and (B) two-dimensional false-color height mapcalculated from the interferogram, taken from Lu¨ttge and Conrad.187

Three different terraces can be identified. At the upper part of thefigure MR-1 has started successfully to establish a biofilm.Reproduced with permission from ref 187 Copyright 2004 AmericanSociety for Microbiology.

Figure 13. Atomic force microscopy image of two MR-1 cellsentrenched in a 104 calcite cleavage face. The cells are entrenchedin a biotically formed pit. The pit is about 130 nm deep, and itswalls are well preserved (Davis and Lu¨ttge, unpublished).

Figure 14. Vertical scanning interferometry image of a trenchformed by a single MR-1 organism in a calcite cleavage face. Themicrobial cell has been removed. The pit is about 2µm long andabout 130 nm deep; note the elevated rim surrounding the pit (Davisand Luttge, unpublished).

362 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 22: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

important general information on nanoparticles in theenvironment; however, with respect to oceanographic set-tings, not much information can be extracted from this work.

Despite this fact, we should note that all carbonate mineralsthat form through either homogeneous or heterogeneousnucleation (as opposed to simple crystal growth by step flow)must experience a nanophase stage during which specificnanophase behavior is likely.200 Specific nanophase behavioris any behavior or property that sets nanoparticles apart fromtheir bulk equivalent, for example, solubility, tensile strength,or hardness. This distinct change in behavior or property isthen used to define a “nanoparticle”.200 Therefore, futureresearch should eventually include this interesting andpotentially very important subject. Nanoparticles are distinctfrom their bulk-phase counterparts for several reasons. Oneimportant fact is that nanoparticles have a large surface tobulk ratio. That means most of a nanoparticle’s moleculesare included in the solid-water interface or at least veryclose to it. Navrotzky201 conducted a very interesting thoughtexperiment based on thermodynamic properties of nanopar-ticles. Navrotzky argues that the observed differences insurface (free) energies of polymorphs that are in thenanophase state may stabilize those polymorphs not stablein the bulk solution. Because a relatively large number ofimportant minerals “show polymorphism with relativelysmall free energy differences between polymorphs, suchcrossovers in stability at the nanoscale may be a rathergeneral phenomenon”.201 If we accept that a nanoparticle stateis a precursor phase for all precipitating minerals, this effectmay be indeed a significant one particularly in low-temperature environments. It could even explain the forma-tion and existence of metastable phases, particularly if inthe bulk or macroregime, the driving force for convertingthe metastable polymorph into a stable one is not too large.

3.6. Calcium Carbonate −Organic MatterInteractions

3.6.1. Overview

Nonthermodynamic behavior of calcium carbonates thatcannot be explained by the influence of inorganic seawaterconstituents is frequently ascribed to the effects of associatedorganic matter. However, organic matter-carbonate associa-tions in natural systems are complex. Only a small compo-nent of dissolved organic material in seawater has beencharacterized at the compound level and the remainingmacromolecular aggregates have generally been classifiedby convention into several molecular weight fractions. Theproblem of an uncharacterized organic matter fraction iscompounded by the manner of its association with carbonategrains. Organic matter can be found occurring on carbonatemineral surfaces via adsorption, precipitation, and copre-cipitation. This may include dead cellular material andsurface-attached living organisms, such as bacteria. Inaddition, organic matter associated with carbonates differsfrom clastic or metal oxide minerals by its association withincarbonate grains as either occluded biogenic material, livingtissue, or organic membranes interspersed between minerallayers.

Investigations of organic matter associated with calciumcarbonate can be divided into two categories. The first andlargest category of study is the behavior of carbonates inthe presence of dissolved organic matter. These investigationsdemonstrate effects that imply indirectly that organic matter

is associated with carbonate surfaces in natural systems. Thesecond and much smaller group of investigations hasexamined the adsorption of organic compounds to carbonatesurfaces under laboratory conditions. Differences in meth-odology in adsorption investigations have led to discrepanciesin the reported affinities of organic compounds for carbonatesurfaces. Processes other than adsorption, such as theinfluence of bacteria at surfaces, precipitation or dissolutionof the carbonate surface during adsorption, coprecipitationof organic material with calcium carbonate, reaction vesselwall adsorption, solubility of organic compounds in seawater,and methods of data analysis, combine to conceal genuineadsorption. Use of nonsterile conditions in most experimentsis a further potential complicating factor.

3.6.2. Indirect Evidence of the Influences of OrganicMatter on Calcium Carbonate

The importance of the interactions of organic matter withcalcium carbonate mineral surfaces in marine systems wasrecognized in the early 1960s. These interactions werestudied primarily because of their effect on carbonate reactionkinetics and, to a lesser extent, the apparent control organicshave on determining crystal morphology.

Surface seawater is supersaturated with respect to calciteand aragonite and thermodynamic calculations indicate thatthis supersaturation should be alleviated by massive inorganicprecipitation. The sediment record shows instead that mostprecipitation of carbonates is biogenic in origin. The absenceof inorganic precipitation has been partially attributed to thepresence of surface-associated organic matter by two possiblemechanisms: (1) organic sheaths, which prevent nucleationor dissolution by physical isolation, and (2) adsorption ofdissolved organic material poisons on sites where activenucleation or dissolution would have occurred.

Kitano and Hood202 carried out a detailed investigationon the influence of organic material on calcium carbonateprecipitation. Using a “homogeneous” precipitation proce-dure, they found that one type of organic compounds favoredthe formation of calcite, while other compounds resulted inaragonite precipitation. Inhibition of precipitation roughlycorrelated with complex formation of the organic compoundCa2+ salt and also increased with increasing organic con-centration. Calcite precipitation was favored by a slowerreaction rate indicating that organic material that inhibitedprecipitation the most also promoted calcite formation.Kitano and Kanamori203 demonstrated that Mg-calcite forma-tion was favored in the presence of the sodium salts of citrateand malate. Investigations by the Kitano group also suggestedthat a site-specific association of the organic material withsurficial Ca2+ could be responsible for inhibition of precipi-tation. Other investigations142,204-207 revealed that carbonateparticles in the water column were coated with organicmaterial forming organic “sheaths” that isolated the normallyreactive surface from seawater and prevented equilibration.Site specific chemisorption or physical coatings were im-plicated as mechanisms for isolation.

Wollast208 and Berner209 confirmed that at the supersatu-ration state of surface seawater, homogeneous precipitationof calcite or aragonite did not occur and that there was noindication of heterogeneous nucleation on abundant sus-pended particulates in shallow-water marine environments.These findings increased the likelihood that surface-associ-ated organic material was responsible for the observedbehavior of carbonates in natural systems. Pytkowicz210

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 363

Page 23: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

concluded that inorganic precipitation of carbonates did notoccur to a large degree due to organic carbon coatings. Heobserved inhibition of calcium carbonate precipitation in thepresence of citrate; however, aragonite formation wasunaffected by stearate. Sodium glycerophosphate was usedas an analog for organic phosphorus in natural waters andinhibition of calcium carbonate precipitation, which wasascribed to adsorption at active growth sites.

Berner211 examined the effect magnesium had on calciteprecipitation and observed that at seawater concentrationsMg2+ retarded calcite precipitation but did not affect aragon-ite precipitation. He concluded that Mg2+ was not adsorbedto the surface of aragonite nor was it found in significant con-centrations within the aragonite lattice. In contrast, dissolvedMg2+ was adsorbed to calcite surfaces and incorporated intoits lattice. Supersaturation with respect to calcite could bepartially credited to the effect of Mg2+ in seawater; however,Mg2+ adsorption cannot resolve aragonite supersaturation.

Berner et al.143 conducted an extensive investigation todetermine inhibition of aragonite precipitation in the presenceof a suite of organic compounds. Short-term laboratoryexperiments using the pH-stat technique to observe precipita-tion rate under steady-state conditions were carried out usinga suite of pure organic compounds and natural fulvate andhumate (Table 2). Aspartate and stearate had little effect onprecipitation rate, but orthophosphate, humate, fulvate, and

several polyfunctional aromatic acids produced large de-creases in precipitation rates. The inhibitory effect ofprecipitation increased with organic and phosphate concen-tration. In addition, rate of precipitation in pore waters wasalso markedly retarded. Berner concluded with short-termlaboratory experiments that aragonite precipitation wassignificantly retarded at low concentrations of dissolvedphosphate and humic matter. However, he stressed that therole of organic matter in inhibiting precipitation is speculationwithout “hard data on the surface chemistry of naturalsuspended particles and the effects of changes in thischemistry upon the ability of the particles to act as seedsfor both aragonite and calcite crystallization”.

Experimental studies have been carried out that empha-sized the role of organic matter in the formation of oo¨ids(see section 4.3.3. for definition and discussion of these sand-sized abiotic carbonates). The primary reason for thisemphasis is the observation of abundant protein matrices inooids.212-215 Mitterer214 noted the abnormally high contentsof aspartic and glutamic amino acids in these proteins andstressed the similarity of this protein composition to that offluids in carbonate-secreting organisms and suggested thatthis proteinaceous matter may play an important role in theprecipitation of aragonite on oo¨id surfaces. The role oforganic matter in the formation of oo¨ids is further discussedin section 4.3.3.

Less work has been done on the effects of organic matteron carbonate dissolution kinetics. Morse49 examined theeffect of dissolved organic matter on calcite dissolution rateusing a number of organic compounds. Two dissolution trialswere conducted using 0.200 and 1.000 mg L-1 of eachorganic component and dissolution rate was reduced by 5%and 25%, respectively. In addition, calcite was soaked for 2weeks in natural seawater to which 10 mg L-1 of eachorganic component was added. A dissolution rate of ap-proximately 90% that of calcite not soaked prior to dissolu-tion, and 80% that of calcite which had been soaked for thesame period of time in organic-free seawater was observed.Reproducibility was approximately(10%, which indicatedthat the observed changes were marginally significant.Morse49 and Sjoberg216 both concluded that dissolution ratewas unaffected in the deep sea by organic matter but maybe affected in pore waters. In addition, they asserted thatnatural inhibition of carbonate dissolution was a function oforganic physical isolation, rather than organic coatingsderived from chemical adsorption.

From the preceding synopsis of organic inhibitory effectson reaction kinetics, it is clear that organics do have aninfluence on carbonate reactivity. However, there is someevidence that surface-associated organic matter may performadditional functions other than that of inhibition of precipita-tion. Hare217,218 and Simkiss219 investigated the protein-mineral matrix found in mollusk shells and suggested thatorganic material acted as a template for calcium carbonatenucleation in addition to possibly determining the crystalmorphology of the mineral overgrowth.

The chief proponent of this hypothesis has been Mittererand co-workers.212-214,220-223 Mitterer and Cunningham223

demonstrated that surface-active organic material that ad-sorbed to calcium carbonate mineral surfaces also complexessome dissolved Ca2+. This would form an environment atthe adsorbed organic-seawater interface having a Ca2+

concentration higher than the bulk solution and permitnucleation of calcium carbonate.

Table 2. Effect of Organic Substances on the SeededPrecipitation of Aragonite from Synthetic Seawater at aSupersaturation of 2.8

organic compoundmax mgof C L-1 R

A. Little or No Inhibition (R g 0.9)amino acids/proteins

aspartic acid 4.2 1.0glutamic acid 3.0 1.0glycine 2.4 0.9arginine 4.8 1.0

aliphatic polycarboxylic acidssodium oxalate 0.6 1.0succinic acid 1.7 0.9disodium EDTA 2.4 1.0

aromatic acidstrimesic acid 2.2 0.9pyromellitic acid 1.8 1.1

B. Weak Inhibition (0.9> R g 0.3)amino acids/proteins

egg albumin 18.0 0.3fatty acidssodium stearate 108.0 0.7

aliphatic polycarboxylic acidsdipotassium tartrate 1.3 0.6potassium H-tartrate 0.5 0.8sodium citrate 2.0 0.3tricarballylic acid 0.7 0.8

aromatic acidspotassium H-phthalate 2.7 0.8gallic acid 0.3 0.6

C. Strong Inhibition (R < 0.2)aromatic acids

gallic acid 0.8 <0.2mellitic acid 1.7 <0.1tannic acid 9.1 <0.2

humic and fulvic substanceshumic acid from marine mud 6.3 0.02fulvic acid from pore water 2.7 0.18

a R is the ratio of precipitation relative to organic free solution.Adapted from ref 142 with the permission of theAmerican Journal ofScience.

364 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 24: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

These investigations are useful because they demonstrateobserved effects on carbonate behavior in seawater andindirectly imply that surface-adsorbed organic matter isresponsible. However, the question that must be raised ishowdo they influence carbonate behavior? Little informationis available about the chemistry of the interfaces that producethese effects. Clearly, there is a need for determining theextent and manner that organics adsorb to carbonate mineralsurfaces.

3.6.3. Direct Evidence for Organic Matter Adsorption onCalcium Carbonate

There are few adsorption investigations quantifying thepartitioning of dissolved organic material in seawater withcarbonate surfaces. Studies have focused primarily on fattyacid and amino acid adsorption over a range of solutioncompositions from distilled water to seawater.

Biogenic carbonates and inorganic carbonates such asoolites are characterized by an organic matrix, which islargely proteinaceous and has a preponderance of acidicamino acids.212,218-219 Jackson and Bischoff224 studied theeffect of amino acids on the calcite to aragonite transforma-tion and the adsorption of free aspartic acid on calcite andaragonite surfaces. Rates of recrystallization were measuredat 66 °C in 0.01 M NaCl solutions in the presence andabsence of individual acidic, neutral, and basic amino acidmonomers. Trials were conducted in the presence of anamino acid and employed 0.01 M solution concentrationsof the respective amino acids. The rate of recrystallizationof aragonite to calcite was accelerated in the presence ofneutral and basic amino acids and retarded in the presenceof acidic amino acids. Inhibition of reaction rate wasattributed to the formation of a protective layer of adsorbedaspartate or glutamate. The extra carboxyl group of aspartateand glutamate, in their dissociated, negatively charged state,theoretically substituted for CO32- ions at the carbonatesurface resulting in chemisorption. Surficial Ca2+ ions wereassumed to be partially chelated by the amino acid moleculesand partially coordinated with CO32- ions in the crystal. Basicand neutral amino acids contained only one carboxyl groupand this was believed to be unavailable for adsorption dueto interaction with an amine group on the amino acid.

The concentration of amino acid used by Jackson andBischoff224 is many orders of magnitude higher than what isfound in surface seawater or pore waters. Extrapolation ofthe results of experiments conducted at high temperature andhigh adsorbate concentration to conditions found in shallow-water marine environments is not necessarily justified. Thevariability of solution pH at which the trials were conductedraises questions about the carbonate saturation state of thesystem and increases the likelihood that processes other thanadsorption prevail. These processes include precipitation ordissolution of the carbonate surface and coprecipitation ofthe amino acid in surface overgrowths. The experimentaltrials were conducted under nonequilibrium conditions withrespect to the carbonate mineral surface and this raises thepossibility that coprecipitation in an overgrowth was respon-sible for uptake of amino acid monomers rather thanadsorption.

Muller and Suess225 found acidic amino acid selectivityover other amino acids in the adsorption of natural organicmaterial to carbonate grains. Despite lack of bacterial controlsand an undefined saturation state of the solution relative tothe carbonate mineral phase, the investigations of the Mitterer

group and Muller and Suess appear to reflect the selectiveacidic amino acid enrichment that is undeniably observedin natural carbonate systems.

The adsorption of fatty acids onto calcium carbonatesurfaces has been less extensively studied. Suess142,206

investigated the adsorption of stearic acid on calcite anddolomite surfaces in distilled water and hexane. Stearate wasadsorbed strongly onto both surfaces in the two solutionmedia examined. Investigations with distilled water indicateda surface saturation monolayer coverage of approximately280 nmol m-2 for calcite and roughly half as much surfacecoverage with dolomite. Adsorption was alleged to occurby chemisorption of the stearate carboxyl in its dissociated,negatively charged form with surficial Ca2+. This mechanismis identical to the site-specific mechanism proposed for acidicamino acid adsorption. It is unclear whether the distilledwater solutions employed in the adsorption trials were inequilibrium with calcite. If they were not, the potential fordissolution of the calcite surface during adsorption ispossible. Dissolution of CaCO3 in distilled water willintroduce Ca2+ and CO3

2- ions, which may reduce thesolubility of stearate and restrict the range over whichadsorption measurements could be conducted. In addition,the potential problems of stearate sorption onto reactionvessel walls and bacterial degradation of stearate were notaddressed.

Lahann and Campbell226 investigated the adsorption ofpalmitic acid on calcite surfaces. In distilled water, adsorptionpartitioning between equilibrium surface concentration andinitial solution concentration yielded an isotherm that was“S” shaped. Adsorption increased up to 25 NaCl-salinity anddecreased thereafter with increasing NaCl-salinity. Decreas-ing solubility with increasing salinity was believed to beresponsible for the increased adsorption up to a salinity of25, and the onset of micelle formation resulted in the decreasein adsorption between salinities of 25 and 35. Magnesiumat concentrations found in seawater at a salinity of 35completely blocked palmitate uptake. Lahann and Campbellhypothesized that palmitate bonded to the calcite surface bysite-specific association with surficial calcium ions (chemi-sorption) and that hydrophobic stabilization was obtained bya parallel arrangement of hydrocarbon chains perpendicularto the mineral surface. Magnesium was believed to inhibituptake by (1) surface exchange of Mg2+ for surficial Ca2+,resulting in a surface with a higher energy of hydration,thereby preventing or severely retarding bond formation, and(2) the formation of a Mg-palmitate complex in solution,reducing the activity of free palmitate available for surfaceadsorption. Phosphate cosorption results were interpreted asindicating that orthophosphate competed for surface sites ona one-to-one basis with phosphate desorption accompanyingpalmitate uptake. Readsorption of phosphate was presumablyblocked by adsorbed palmitate molecules. Bacterial metabo-lism of adsorbate, reaction vessel wall adsorption, thesaturation state of the solution with respect to calcite, andpalmitate solubility were not considered in this study. Inaddition, isotherms were reported using initial solutionconcentration.

The adsorption behavior of several normal saturated fattyacid homologs, twelve free amino acids, and three poly-glutamate molecules of different monomer unit size on thecarbonate mineral surface in seawater and related solutionshave been described and quantified.144,227 Synthetic poly-glutamic acids were employed to model acidic amino acid-

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 365

Page 25: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

rich proteins. Laboratory investigations required extensiveprocedural modifications from those of previous investigatorsto ensure that observed partitioning was a function of sorptionprocesses. Sorption behavior of fatty acids adhered toTraube’s rule and was complex as indicated by S-shapedisotherms. Free amino acids did not sorb, but polyglutamicacids did.

The classes of organic compounds investigated contain thecarboxyl function. The carboxyl group has been widelymentioned as the function responsible for the adsorption offatty acids, amino acids, fulvates, and humates onto CaCO3.Although no direct proof of a mechanism for sorption ofthis nature has been observed, sorption was hypothesized tooccur by ligand exchange with surficial Ca2+. Results of theZullig and Morse144 investigation indicate that sorption isfar more complex and the presence of a carboxyl functionand surficial Ca2+ does not necessarily ensure adsorption ontoCaCO3 surfaces. For example, at low solution concentrationsconsistent with levels reported in seawater, shorter chain fattyacids (C1 to C12) and all of the free amino acids, includingthe acidic amino acids, did not adsorb onto CaCO3 surfaces.Yet, they all contain the carboxyl function. Sorption of C14

to C18 fatty acid homologs onto dolomite and magnesitesurfaces at low solution concentrations is similar to sorptiononto calcite and aragonite surfaces. This behavior is difficultto explain as a function of a single site-specific mechanism,even if surface sorption site heterogeneities are considered.Free glutamate did not adsorb, but the peptide-bondedpolymer of glutamate sorbed onto calcite surfaces.

In summary, the interactions of different organic com-pounds and their influences on calcite and aragonite reactionkinetics varies widely. Observations of carbonate mineraldiagenesis in marine sediments indicates that biologicallyderived organic matter may have little influence on calciumcarbonate dissolution but may play a major role in inhibitingprecipitation on sedimentary carbonates.

4. Oceanic Sources of Marine Carbonates

4.1. OverviewAlthough near-surface seawater is strongly supersaturated

with respect to calcite (∼6×), aragonite (∼4×), and idealdolomite (∼50×), these minerals fail to precipitate directlyfrom it under near-normal oceanic conditions (see sections2 and 3). In the modern ocean, the strongly dominant sourceof carbonate minerals comes from a wide variety of organ-isms that yield “skeletal” carbonates. In the pelagic realmof the open ocean, the primary carbonate mineral formed islow-Mg calcite (<4 mol % MgCO3), with smaller amountsof aragonite also being derived from organisms. Deep oceansediments contain high-Mg calcite (>4 mol % MgCO3) frombenthic organisms and occasional abiotic marine cementsusually as a very minor component of total sedimentarycarbonate minerals.

In the shoal-to-shallow water depth of temperate-to-tropicalregions of the ocean, biogenic aragonite is the most abundantcarbonate mineral, followed by high-Mg calcite and then low-Mg calcite. In cold water environments, typical of highlatitudes, low-Mg calcite is dominant. Dolomite is relativelyrare and largely confined to “special” environments. Al-though in a great many instances, it is clear that thesedimentary carbonates are the skeletal remains of organisms,considerable controversy exists as to the role of organismsin forming fine-grained aragonite muds, marine cements, and

more recently dolomite. In siliciclastic sediments Fe(II)- andMn(II)-containing carbonate minerals and massive or nodularcalcite and aragonite are the result of processes taking placein sediments that involve organisms but are abiotic productsof these biologically driven reactions.

It is important to keep in mind that the sources, depositionand, diagenesis of marine carbonate minerals to a large extentreflect the biology of the ocean at least as much as itschemistry. This can lead to circumstances where dissolutionof carbonate minerals in apparently supersaturated watersand their deposition from apparently undersaturated waterscan occur because of local biologically driven environmentsof altered saturation state. Furthermore, most oceanic carbon-ate minerals may be repeatedly passed through the guts oforganisms feeding on suspended matter or near-surfacesediments. In recent years, the role of organisms has beenargued to be important even for carbonates that havetraditionally been considered abiotic (e.g., dolomite) throughthe role of highly controversial “nanobacteria” that are muchbeloved in sections of the Earth science community anddenounced as impossible by many microbiologists.

Considerable effort has gone into modeling the cyclingof carbonates in the ocean. Although biogenic formation ofcalcium carbonate is domoninantly in the open ocean basins,the accumulation of calcium carbonate in sediments is greaterin the coastal oceans.228 Most commonly oceanic carbonatesare considered as part of the broader general oceanic orglobal carbon cycles. These can be quite complex such asthe recent shallow-water ocean carbonate model (SCOM).3,229

In this paper, we will not go into a discussion of these typesof models. For those interested in pursuing this area further,a good start would be the recent review by Sabine et al.230

“Current status and past trends of the global carbon cycle”.

4.2. Primary Carbonate Sources for Deep SeaSediments

The primary sources of calcium carbonate for deep seasediments are pelagic calcitic coccolithophores and fora-minifera that grow dominantly in the phototic zone (typicallyupper∼100 m) and to a lesser extent aragonitic pteropods.The distribution of calcium carbonate-secreting pelagicorganisms is primarily controlled by the fertility and tem-perature of the near-surface ocean. The fertility of seawateris largely a result of ocean circulation patterns and, inparticular, those processes leading to upwelling of nutrient-rich waters.

In general, coccolithophores are common in temperatewaters, but rare in high latitude cold waters where diatoms(siliceous skeletons) dominate. Coccolithophores are numeri-cally much more abundant (usually∼10 000-fold) thanpelagic foraminifera. Benthic foraminifera can be formed ofeither aragonitic or high-Mg calcite. Because of their generalscarcity in deep sea sediments, aragonitic pelagic organismshave received relatively little attention. Pteropods andheteropods are the most abundant pelagic aragonitic organ-isms. It is estimated231 that the percentage of the CaCO3 fluxis about 12% of the calcium carbonate flux from near-surfacewaters, although this percentage may vary considerably indifferent areas of the ocean and seasonally.

The mechanisms and rates of transfer of biogenic carbon-ate material from near-surface waters to deep sea sedimentshave been intensely investigated. Major questions have beenconcerned with the transition from living organism to carbon-ate test, rates of sinking, extent of dissolution in the water

366 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 26: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

column and on the sea bottom (dissolution is discussed in sec-tion 5), and the relation between life and death assemblages.

Because of their small size, individual coccoliths sinkslowly232,233and, consequently, may spend long periods oftime (on the order of 100 years) in the water column. Thislong residence time should lead to substantial to completedissolution of the coccoliths in the undersaturated part ofthe water column. The origin of coccolith ooze on portionsof the seafloor overlain by undersaturated waters, therefore,is difficult to explain by considerations based on settling ratesof individual particles.234,235 Data obtained from sedimenttraps has shown that most coccoliths reaching the seafloordo so by aggregation caused by zooplankton grazing andencapsulation in fecal matter.

Submarine lithification and precipitation of cements indeep sea carbonate sediments are relatively rare processesin typical major ocean basin sediments.235-238 The cementsare of both aragonitic and magnesian calcite mineralogies,and are largely restricted to shallow seas such as theMediterranean and Red seas, and sediments in the shallowerparts of major ocean basins in which biogenic aragonite isalso present. Many shallow water and deep sea carbonatecements fall in the range of 10-15 mol % MgCO3 with astrong maximum in abundance at about 13 mol % Mg.21,239

Another often local source of aragonite and high-Mgcalcite to the open ocean water column and deep seasediments is shallow-water carbonate-rich sediments. Duringevents such as major storms sediments in the shallow-waterenvironments can be resuspended and transported to nearbydeep-water environments. Dissolution can occur in the upperwater column, as demonstrated near Hawaii for high-Mgcalcite240 or be deposited in deep sea sediments (e.g., nearthe Bahama Islands241,242).

4.3. Shallow Water Biogenic Carbonate Sources

4.3.1. General Considerations

As previously described, the grains of carbonate sedimentscan be subdivided into skeletal and nonskeletal components.Carbonate muds, composed dominantly of silt- and clay-sized (<63µm) particles, are common deposits in low-energyenvironments, such as tidal flats and subtidal areas. Largercarbonate grains, which may dominate to form carbonatesands, include pellets that are formed by the ingestion ofsediment by marine organisms and excretion of fecalmaterial; oo¨ids, which are spherical to ovoid 0.2-1 mmgrains with an internal concentric or radial structure; litho-clasts, which are fragments of previously deposited andusually somewhat lithified carbonate sediment; and relictgrains of older origin, having formed under previousenvironmental conditions.

The skeletal components of carbonate sediments representthe complete or partial skeleton or the decomposed anddisaggregated skeletal remains of organisms extant at thetime of deposition of the sediment. Sedimentary particlescan be produced from calcareous organisms by disaggrega-tion of their skeletons, by mechanical means related to waveand current energy, or by bioerosion of carbonate substrateslike corals, mollusks, and rocks. The wide diversity in thetypes of shoal-shallow water organisms and their complexcompositional and morphologic characteristics, in combina-tion with the varied processes causing their disaggregation,lead to a very complex mixture of biogenic carbonate grainsin shoal-shallow water sediments.

4.3.2. Sources of Fine-Grained Carbonate Sediments

The origin of carbonate mud is a long-standing topic ofconsiderable debate and controversy. The three majorpotential sources generally considered are (1) physiochemicalor biochemical precipitation, (2) postmortem disintegrationof calcareous algae, and (3) physical or biological abrasionof skeletal material.243 Research has generally focused oneither aragonite or high-Mg calcite mud components. Gener-ally within the silt and clay sizes (<63 µm), there is littlemineralogic, elemental, or stable isotopic variation amongsubdivisions of this grain-size fraction.244

Fine-grained carbonate-rich sediments that are dominatedby small (a few micrometers) needle-shaped aragonite grainsare referred to as aragonite needle muds. The two majorsources for the aragonite needles generally considered arecalcareous algae and abiotic precipitation.245,246The fine sizeof this material has resulted in most of the interpretation ofits origin being largely based on chemical data, althoughneedle morphology has also been used.

Attempts at making budgets of carbonate production andaccumulation have also been used to attempt to answer thequestion of the source of aragonite needle muds. In areassuch as British Honduras and Florida Bay, the supply ofbiogenic material appears to be sufficient to provide thesediment for the carbonate muds, although some abioticprecipitation cannot be ruled out. However, in other areassuch as the Great Bahama Bank and the Persian Gulf, thebiogenic supply appears to be insufficient, and an abioticsource of aragonite needles is needed.247 A basic difficultyin determining the origin of aragonite needles is thatprecipitation of aragonite from seawater produces needleswith a very similar size and morphology to those resultingfrom breakdown of common codiacean green algae. It isnecessary, therefore, to examine the chemistries of theneedles to determine their origin. However, even the chemi-cal characteristics have generally not proven unambiguouslydiagnostic.244,245,248,249

The waters and sediments of the Great Bahama Bank havebeen intensely studied by carbonate geochemists. Followingis a brief summary of some of the more recent findings thatprovide an excellent opportunity to tie laboratory results tofield observations.250 The study was aimed at understanding“whitings”, which are suspended “clouds” of fine-grainedCaCO3 that can even be observed from space (Figure 15).They occur on the northern part of the Great Bahama Bank(GBB) and on the Little Bahama Bank (LBB). Their abioticversus biotic origins have been widely disputed. On the GBB,the waters have a long residence time (∼240 days), and abouta third of the specific alkalinity is lost due to CaCO3

precipitation.The “kinetic solubility” of suspended whiting material has

been determined to be about 1.9 times that of aragonite andinner Bank waters are close to equilibrium with the phasethat gives this kinetic solubility.250 Similar values of the ionconcentration products were observed for many pore watersfrom carbonate-rich sediments in this region.251 The kineticsolubility value obtained for the dominantly aragoniticsuspended sediment from the Bahamas was very similar tothat obtained for thin overgrowths on calcite.250

The rate of calcium carbonate precipitation in waters onthe GBB was calculated from changes in specific alkalinity,estimated water mass ages and the amount of suspendedcarbonate solid in the waters. The rates were in goodagreement with experimentally determined rates using sea-

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 367

Page 27: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

water at similar saturation states and a fine fraction ofsediments obtained from the area where the whitingsoccurred. These results are encouraging in demonstrating thatsolubilities and precipitation kinetics obtained from both thefield observations and laboratory experiments were in goodagreement.

Although not nearly as much attention has been paid tothe production of lime mud from calcitic sources as aragonite,it can be the dominant component in perireefal251 andlagoonal237 environments where Mg-calcite may constitutefrom 70% to 90% of the mud. The mud can form frommicritization and recrystallization of skeletal grains.237

Breakdown of foraminiferal tests can often be a primarysource of Mg-calcite in muds.252

4.3.3. Sources of Carbonate SandsCarbonate-rich sediments that are dominated by sand-sized

(>63µm) grains are common in shoal-water sediments. Theyare usually associated with moderate- to high-energy envi-ronments such as beaches, edges of carbonate banks, andbehind fringing reefs. Carbonate sands are often composedof carbonates of mixed mineralogy from a variety of biogenicsources (Table 3). These components are produced byphysical abrasion, boring, and other biological processes.

Subspherical carbonate grains, generally referred to asooids because of their resemblance to fish eggs, comprisesandy sediments called oo¨lites that are common membersof modern shoal-water tropic and subtropic carbonate-richsediments. Oo¨ids are variable in both mineralogy andstructure. Aragonite, high magnesian calcite, and calcite havebeen observed to occur in oo¨ids.253 Radial, concentric-tangential (laminated), and micritic structures are commonlyobserved around a central nucleus of nonoo¨id material. Oo¨idsare believed to be of nonbiogenic origin and represent oneof the most important modes of nonbiogenic removal ofCaCO3 from the ocean. In addition, it has frequently beenpointed out that their mineralogy and structure reflect thoseof marine carbonate cements.253,254 In fact, they grade intoeach other in the case of Bahamian grapestone. Because oftheir importance as presumably abiotic marine carbonates,ooids have been extensively studied and brief summary ofthe primary findings follows.

There are three primary classes of hypotheses for theprocesses responsible for the existence of oo¨ids. They arebased on bacterial-mechanical, algal, or chemical mecha-nisms for oo¨id formation. Some investigators have madehybrid models involving more than one of these mechanisms.We will only present these hypotheses in a general manner

Figure 15. NASA space shuttle photograph of whitings (filamentous white linear features near center) taken of the central Little BahamaBank (Roll 719; Frame 29). Image courtesy of the Image Science & Analysis Laboratory, NASA Johnson Space Center (http://eol.jsc.nasa.gov).

368 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 28: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

and not attempt to reference the large number of papersrelevant to the topic.

One of the earliest nonbiological hypotheses255 for themechanism of oo¨id formation was that oo¨ids with aconcentric structure form by a process similar to that whichforms a snowball. In this model, an organic coating isbiologically deposited (the infamous “mucilagenous slime”)on the surface of the oo¨id and then, as the oo¨id is rolledback and forth by wave action, small aragonite needles,which are common in most modern depositional environ-ments where oo¨ids are found, become imbedded in theorganic matter. During subsequent quiescent times a neworganic coating is added. This process is repeated many times(up to 200 lamellae have been observed in oo¨ids) causinggrowth of the oo¨id. This model has the attractive character-istics of not requiring biotic or chemical formation of CaCO3

directly on ooid surfaces and offering an explanation for theobservation that oo¨ids appear to require “rest” betweendepositional cycles. It also provides a reasonable explanationfor the concentric organic-rich layers generally observed inlaminated aragonitic oo¨ids.

There are, however, some problems with the model.1 Itclearly does not apply to radial oo¨ids. When examinedcarefully, it fails to explain the general excellent sorting ingrain size and mineralogy, tangential arrangement of thearagonite rods, or the formation of relatively thick lamellaeon typical aragonitic oo¨ids. It is also questionable how wellsuch coatings could hold up on grain surfaces in the abrasive,high-energy hydrodynamic environments characteristic of

laminated aragonitic oo¨id formation. It may be best toconsider this bacterial-mechanical hypothesis for the forma-tion of laminated aragonitic oo¨ids as a precursor to morerecent models, where adsorption of organic matter on oo¨idsurfaces influences the precipitation of CaCO3.

The hypothesis that oo¨ids form by direct chemicalprocesses has gained increasing popularity, and there is agrowing body of experimental and observational evidenceto support it. Stable isotope240-245 and Sr concentration dataare generally compatible with nonbiogenic formation. Inrecent years, particular emphasis has been placed on thepossible role of organic compounds.256,257 This hypothesisis also not without its problems. It fails to explain why ithas been difficult to produce nonradial oo¨ids by directprecipitation, it fails to give an explanation for organiclamellae or the general distribution of oo¨ids, and it cannotexplain why high magnesian calcites are formed in calciticooids.

Bathurst258 has presented a model that solves some of theproblems associated with laminated aragonitic oo¨id forma-tion. His suggestion involves a means for growing inorganicprecipitates of tangentially oriented rods. His arguments alsooffer plausible explanations for limitations on oo¨id size andother characteristics. One of the difficulties in understandingtypical aragonitic oo¨id formation is the fact that the aragoniteneedles, or battens, are arranged with theirc crystallographicaxes parallel to the solid surface. Bathurst258 observed thatthe radial-fibrous fabric found in calcitic oo¨ids can beexplained as the result of competition between crystals thatgrow most rapidly in a direction parallel to one crystal-lographic axis. The perpendicular orientation of the aragonitec-axis is commonly observed for aragonite crystals on thewalls of Foraminiferida and in the utricles ofHalimeda. Itis also the preferred orientation of aragonite in beach rock.However, it is not the typical orientation of aragonite on freecalcrete grains where the aragonite grains are arranged withtheir c axes parallel to the surface. Bathurst259 observed thegrowth of aragonite needles with theirc axes parallel to thesurface in 3µm thick films on a variety of calcarenite grainsoutside of the oo¨litic shoal environment in the coralgal, stablesand, grapestone, and mud habitats at a number of locations.He suggested that the tangential orientation may result fromthe adsorption of colloidal sized aragonite nuclei on the solidsurface, with their longest axis parallel the surface providingthe most stable arrangement.

Contrary to the contention of many investigators of oo¨idformation, Bathurst suggests that there is not a strongcorrelation between the degree of water agitation and theorientation of aragonite needles. He cites his observation ofthe similarity between grains with oo¨litically structured filmsin the Bimini Lagoon and the oo¨ids of Brown’s Cay,Bahamas, which exhibit no petrographic differences, includ-ing the existence of lamellae, in spite of the radically differenthydrographic environments. His ideas have been reinforcedby the observations of Land et al.253 in Baffin Bay, Texas,where calcitic radial-fibrous and aragonitic tangential oo¨idscoexist in the same sands and show no significant correlationwith the probable wave energy differences at the differentsampling sites.

Bathurst259 also examined the question of size and sortingof ooids and the thickness of individual oo¨litic coats. Hefeels that the thickness of the coats may reflect growth rates,which are in turn a reflection of the degree of agitation.Higher energy waters will expose the oo¨ids to more

Table 3. Skeletal Composition of a Calcareous Sand from aGrass Bed, Ferry Reach, Bermudaa

taxa identified number of grains percent

gastropods 162 6.4bivalves 81 3.2foraminifera 722 28.7grass fragments 205 8.2Halimedaspp. 1117 44.5Cymopoliaspp. 26 1.0ostracods 36 1.4encrusting red algae 9 0.4serpulid (?) worm tubes 9 0.4agglutinate worm tubes 11 0.4caecum gastropods 10 0.4unidentified grains 90 3.6total grains countedb 2515 98.6

type of organism number of species

gastropods 9bivalves 5foraminifera 21grasses 3Halimedaspp. 3Cymopoliaspp. 1other green algae 1encrusting red algae 1chitons 1ostracods 3echinoids 1agglutinated worm tubes 1crabs 1serpulid worms 1scaphopods 1caecum gastropods 1sponges 1total species counted 55

a Adapted from ref 1.b The remainder included non-calcareous greenalgal fragments, chitons, echinoid spines, scaphopods, sponge spicules,and crab carapace fragments.

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 369

Page 29: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

supersaturated water for a longer period of time. A problemwith the sorting and size of oo¨ids on banks is that the oo¨idsare usually of similar size. Small ones should be winnowedfrom the oolite banks by the intense hydrodynamic conditionsgenerally present. The source of the oo¨ids, therefore, mustbe in deeper waters where hydrodynamic conditions are lessintense. Bathurst suggested that the upper size limit on oo¨idsmay be governed by a balance between abrasion and growthrates.

Weyl260 observed carbonate precipitation in seawaterflowing over a column of oo¨ids. After an initial period ofprecipitation, the reaction between the oo¨ids and supersatu-rated seawater appeared to stop. If the oo¨ids were allowedto sit for a period of time with no water flowing, they wouldagain become reactive. This experimental observation ledto the idea that oo¨ids need to “rest” between reactive periods.It was hypothesized that a high magnesian calcite coatinginitially formed, which later dissolved in pore waters of anoolite bank before more aragonite would precipitate. Thishypothesis has been quite popular ever since Weyl’s work,although it is far from proven.

By far the most extensive experimental study of thechemistry of oo¨id formation was done by Davies, Bubela,and Ferguson.261,262 Their primary emphasis was on theinfluence of different types of organic compounds andagitation conditions on the mineralogy and morphology ofthe precipitated carbonates. After studying the influences ofa variety of organic compounds on the mineralogy andmorphology of the precipitates, it was found that humic acidswere most effective in producing magnesian calcite (com-position was not reported), with a fabric similar to that ofradial ooids. It was also observed that, while low molecularweight humic acids were preferentially adsorbed relative tohigh molecular weight humic acids, the high molecularweight humic acids were most effective in producing oo¨id-like products. A lengthy discussion of the possible reasonsfor this observation was presented, which noted that the highmolecular weight humic acids had a higher Mg to Ca ratioand that humic acids through hydrophilic-hydrophobicinteractions could possibly make membrane-type structures.Tangentially oriented batten-like aragonite crystals weresynthesized on the surfaces of quartz grains and Bahamianooids without the intervention of organic compounds. A highdegree of agitation was found necessary, and like Weyl,Davies, Bubela and Ferguson observed the necessity of a“resting” period during which surfaces were reactivated,possibly by the release of Mg2+ or H+. Their studies areremarkable in indicating that organic matter may be impor-tant for the formation of magnesian calcite radial oo¨ids butnot aragonitic tangential oo¨ids. This observation is contraryto general consensus on the formation of these two types ofooids. It may also offer a major clue to the formation ofaragonitic and calcitic carbonate cements.

4.3.4. Carbonate Cements

The formation of carbonate overgrowths and cements iscertainly one of the most important and highly studied aspectsof the behavior of carbonates in sediments and sedimentaryrocks1,263-265 Carbonate cements in calcareous sediments fallinto three major groups. The most common are those thatoccur in voids found in biogenic carbonates. This group isespecially important in reefs. Cements also occur on theexterior of carbonate particles where their intergrowth cancause formation of hardened pellets, grapestones, crusts,

hardgrounds, and beachrock. This type of cementation mayalso be important in oo¨ids.254 Micritic cements associatedwith boring algae comprise the third common type of cement.

One of the most interesting questions about carbonatecements in the marine environment is why they are not moreabundant. In carbonate-rich sediments, there are abundantcarbonate surfaces for overgrowths to form on, and bothoverlying seawater and many pore waters are supersaturatedwith respect to carbonate minerals. Morse and Mucci266

demonstrated that precipitation of cements from pore watersof carbonate sediments is severely inhibited by buryingIceland spar calcite crystals in a variety of carbonatesediments in the Bahamas for several months. After recoveryand analysis using depth-profiling Auger electron spectros-copy, they were able to detect overgrowths on only a fewcrystals. The calculated growth rates of the precipitates weremuch less than those observed in laboratory experimentsusing seawater at the same supersaturation state.

At least part of the reason for these observations must bethat precipitation is probably severely inhibited by organicssuch as humic acids.267 Thus, inhibition of precipitation,coupled with slow transfer of fresh supersaturated seawaterinto sediment pores, seems to account for lack of extensiveearly cementation.

Factors controlling the mineralogy and chemical composi-tion of carbonate cements in the marine environment havealso been extensively investigated. Bathurst268 summarizedmany of the observations269 indicating that in some casesthe host carbonate mineral is a factor in determiningcomposition. Usually aragonite grains have aragonite over-growths, whereas high-Mg calcite grains have high-Mgcalcite overgrowths of similar Mg content. In some instances,syntaxial formation of cement is observed. Although thisexplanation of host control is a “comfortable” one, it doesnot explain the common occurrence of cements of mixedmineralogy or those that differ from their host grains.1

A major concern in studies of the chemistry of carbonatecement formation is the source of the calcium and carbonateions necessary to form the cements. The obvious source isseawater, but large volumes of seawater are necessary ifsignificant amounts of cement are to be produced. Cementformation is consequently favored near the sediment-waterinterface and in high-energy environments where water canbe flushed through porous structures such as reefs. Theobservation that cements usually form only in thin crustsnear the sediment-water interface also demonstrates theimportance of normal seawater for cement precipitation.Further strong evidence for cement formation from normalseawater comes from stable isotopes. Theδ13C values ofcements are usually close to those predicted for carbonatesprecipitating from seawater.270 Another possible source ofthe ions necessary for cement formation is from the dissolu-tion of carbonate phases more soluble than the cements.High-Mg calcites could provide such a source, as has beendemonstrated in periplatform oozes in the Bahamas271 andfor coastal carbonate sediments of the Bay of Calvi inCorsica.272 This process, leading to cement precipitation, isconsistent with the previously discussed chemistry of manypore waters.

The precipitation of calcium carbonate usually, but notexclusively, from petrographic evidence in the form of high-Mg calcite273 is generally inferred from a decrease indissolved pore water calcium.274,275 Most studies of theimpact of chemical diagenesis on the carbonate chemistry

370 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 30: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

of anoxic sediments have focused primarily on the fact thatsulfate reduction results in the production of alkalinity, whichcan cause precipitation of carbonate minerals.209 Carbonateprecipitation can also occur via methane oxidation insteadof organic matter oxidation, resulting in distinctly “light”δ13C values for authigenic carbonates.276

Finally, it should be noted that while most investigatorshave rejected the idea that carbonate cements are directly oreven closely linked to biological mechanisms of formation,some investigators hold the position that organisms may bevery important to cement formation processes. Marinepeloids (aggregates of cemented carbonate grains) havereceived considerable attention in this regard because of theirclose association with bacterial clumps.277 In addition,Fabricius254 has presented a lengthy discussion in favor ofthe formation of grapestone cements and oo¨ids by algae.

4.3.5. Dolomite Formation

The formation of dolomite under conditions encounteredin modern marine sediments and sediment burial to moderatedepths is strongly controlled by reaction kinetics that are sloweven at high supersaturations.125,278A plausible explanationfor the slow precipitation kinetics of dolomite is that therequirement of cation ordering puts a major limit on the rateat which the phase can form.279 It is likely that thedehydration kinetics of Mg2+ also play a role.245 Becausethe precipitation kinetics of dolomite are too slow to bestudied in the laboratory at near Earth-surface tempera-tures,280 experiments on dolomite reaction kinetics havegenerally been conducted at elevated temperatures (typicallybetween 100 and 300°C). While considerable interestinginformation has been gathered from these experiments, theapplicability of the results to lower temperature and slowerreaction kinetics remains highly questionable.

Observations of dolomite formation in natural systems alsohave been used as the basis for discerning additional factorsthat may influence the rate of dolomite formation. Theseinclude catalysis by certain clay minerals281 and productionof organic byproducts by bacteria.282 Mg2+ transport to sitesof dolomite precipitation can inhibit the reaction in hemi-pelagic sediments.283 However, the true influence of reactionrates is largely speculative because the kinetic factors aregenerally deduced primarily from the presence or absenceof dolomite in different environments.

Most modern dolomite is forming from high ionic strengthsolutions that are typically derived from the evaporation ofseawater or lakes in arid regions. These environments havebeen studied extensively because they provide an opportunityto observe directly systems in which substantial amounts ofdolomite are currently forming. The associated hydrologyand solution chemistry of the dolomitizing fluids can alsobe determined. Sites of particular note are the Persian Gulfsabkhas, Coorong district in South Australia, Bonaire Islandin the Caribbean Sea, Sugarloaf Key in Florida, supratidalsediments on Andros Island in the Bahamas, and Deep SpringLake in California. A general feature of many of these areasis that high Mg2+ to Ca2+ solution ratios appear to promotedolomitization. Friedman284 discussed the general importanceof modern high salinity environments for dolomite formationand evidence in the rock record indicating a strong relationbetween evaporites and dolomite. He claimed that mostsedimentary dolomite should be considered an evaporitemineral- an opinion that is not universally accepted.

The importance of inhibition of dolomite formation bysulfate has been widely debated.285 While dolomite formationis usually associated with sediments where sulfate reductionis active, the observation that dissolved sulfate concentrationscan remain high and dolomite still can form argues againstdominant control by dissolved sulfate. This observationsuggests that reaction products of sulfate reduction, such asincreased alkalinity, may play a more important role inincreasing the formation rate of dolomite than sulfate doesin inhibiting it. This conclusion has been expanded upon byCompton,286 who emphasized the general observation thathighly elevated solution supersaturations with respect todolomite are usually associated with sediments containingabove average concentrations of organic matter. He termeddolomite that forms under these conditions “organogenicdolomite.” More recently, Mazzullo287 has further exploredand summarized the data supporting the early formation ofdolomite in organic-rich sediments where extensive sulfatereduction and methanogenesis occur.

Arvidson and Mackenzie288,289conducted experiments ondolomite growth at elevated temperatures from which theywere able to derive a rate equation where temperature andsaturation state were used to predict growth rates. Thisequation was used to extrapolate to conditions wheredolomite formation is currently occurring in sediments, andreasonable (within an order of magnitude) agreement betweencalculated and observed growth rates was found.

Another factor that may promote dolomite formationduring early diagenesis is direct participation by bacteria.The literature relating to this possible avenue of lowtemperature dolomite formation was reviewed by Vascon-celos and McKenzie290 in their study of microbial mediationof dolomite formation in Lagoa Vermelha in Brazil.

4.3.6. Carbonate Minerals Formed in SiliciclasticSediments

Carbonate minerals that are formed in siliciclastic sedi-ments are dominantly the result of diagenetic reactions takingplace in the sediments. The quantitatively most importantreactions generally occur very close to or within less than ameter of the sediment-water interface. These reactions areoften associated directly or indirectly with a variety orterminal electron acceptors, such as solid iron and manganeseoxides and dissolved sulfate, used by bacteria metabolizingsedimentary organic matter or hydrocarbon gases. Not onlydo bacteria play an important role in the formation ofcarbonate minerals in sediments, but they are also, as willbe discussed in section 5, the primary driving force fordissolution of sedimentary carbonate minerals. This is a largeand complex field of study that we will only briefly touchon in this paper. However, reviews dealing specifically withthe topic are available.1,291,292

A very important reaction for both the formation anddestruction of calcium carbonate in marine sediments isbacterial sulfate reduction, which can be represented as anapproximation for marine organic matter as given in eq 4.1.

Because the components of the organic matter are at constantratios, fixed pH can be established after a moderate degree

153

(CH2O)106(NH3)16H3PO4 + SO42- f CO2 +

HCO3- + HS- + 16

53NH3 + 1

53H3PO4 + H2O (4.1)

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 371

Page 31: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

of sulfate reduction in a closed system.293 During these earlystages of reaction in seawater, the pH drops to about 6.9, inthe absence of other reactions, and remains constant. Thiscan result in pore waters becoming undersaturated andcarbonate dissolution occurring. Further sulfate reductioncauses the alkalinity to rise while the pH remains constant.At about 35% sulfate reduction, the pore waters regainsupersaturation with respect to calcium carbonate and furthersulfate reduction can result in calcium carbonate precipitation.

Massive carbonate crusts and nodules can be formed insediments rich in organic matter. Sediment through whichhydrocarbon gases, usually dominantly methane, are seepingare especially noteworthy in this regard and are commonlyassociated with deeper sedimentary methane hydrates.294-296

“Reactive” forms of Fe(III) and Mn(IV) are found insiliciclastic sediments that can be reduced either by bacteriaor through reactions with dissolved reduced pore watercomponents such as H2S. The resulting dissolved Fe(II) andMn(II) can react with CO32- ions to either coprecipitate withcalcium carbonate minerals or form distinct iron andmanganese carbonate minerals. For Fe(II) the commonmineral formed is siderite (FeCO3), which is most commonin estuarine and freshwater sediments where little sulfide isformed that competes for the Fe(II) to form minerals suchas pyrite (FeS2). Although Mn(II) can form simple MnCO3(rhodochrosite), careful studies10 indicate that it is more likelyto form a poorly ordered Mn analogue to dolomite, (pseudo)-kutnahorite (CaMn(CO3)2).

5. Dissolution of Calcium Carbonate

5.1. Dissolution of Carbonate Minerals in thePelagic Environment

5.1.1. General Considerations

One of the most controversial areas of carbonate geochem-istry has been the relation between calcium carbonateaccumulation in deep sea sediments297 and the saturation stateof the overlying water (see Millero2 in this issue). Theaccumulation of calcium carbonate in deep sea sedimentshas been mapped for over 100 years based on the “carbonatecompensation depth” (CCD). Although this has been definedusing various concentrations of CaCO3 in sediments, itbasically divides deep sea sediments into those with little(less than∼10 wt %) to no CaCO3 from those containingsubstantial CaCO3 concentrations that commonly exceed 80wt %. Analogous to the CCD is the aragonite compensationdepth (ACD), which is defined in reference to the mineralaragonite rather than simply total CaCO3 content as is theCCD. Two other commonly used sediment markers ofcarbonate dissolution based on foraminifera tests are theforaminiferal lysocline (FL) and the R0 depths. The CCD,FL, R0, and ACD have been carefully mapped in many areas.However, with the exception of complete CaCO3 dissolutionat the CCD and complete aragonite dissolution at the ACD,the extent of dissolution that has been observed in most deepsea sediments is difficult to quantitfy. Consequently, it isgenerally not possible to make reasonably precise plots ofpercent dissolution versus depth. In addition, the analyticalchemistry of the carbonate system and constants used tocalculate the saturation states of seawater have been a sourceof contention. A reason that there has been so muchcontroversy associated with the relation between the extentof carbonate dissolution occurring in deep sea sediments and

the saturation state of the overlying water is that models forthe processes controlling carbonate deposition depend stronglyon this relation. Hypotheses have ranged from a nearly“thermodynamic” ocean where the CCD and ACD are closeto co-incident with calcite and aragonite saturation levels,298

to a strongly kinetically controlled system299 where majordifferences in the CCD, FL, and saturation depth exist.However, it now appears that the long cherished idea of a“tight” relation between seawater chemistry and carbonatedepositional facies must be reconsidered.300 This is a fertilefield for further research.

5.1.2. Dissolution in the Water Column

Millero2 (this issue), discussed that, based on changes inwater chemistry, it is estimated that probably 40-80% ofthe CaCO3 formed in the pelagic environment dissolves inthe upper water column. Maximum dissolution rates occurnear the aragonite saturation depth.228 An extensive inves-tigation of planktonic foraminifera in sediment trap samplesfrom the tropical Pacific and central Atlantic oceans has beenconducted, and at both study sites, the total foraminiferalflux and the carbonate flux tend to decrease with depth.301

In addition, the flux of individual species of planktonicforaminifera varies significantly with depth, with the numberof small, solution-susceptible species decreasing with in-creasing depth. These results indicate that there is significantdissolution of small (<150 µm) foraminifera as they settlethrough the water column. Material collected from thesediment-water interface directly below the Pacific sedimenttrap array contained no planktonic foraminifera, suggestingthat the residence time of an individual skeleton on theseafloor before it dissolves is extremely short.301,302

A study of the sedimentation of pteropods and foraminiferain the North Pacific using sediment traps confirmed thatconsiderable dissolution of pteropods was taking place inthe water column.303 It was calculated that approximately90% of the aragonite flux was remineralized in the upper2.2 km of the water column. It should be noted that the depthfor total dissolution in the water column is considerablydeeper than the aragonite compensation depth. This isprobably due to the short residence time of pteropods in thewater column because of their rapid rates of sinking.

5.1.3. Dissolution on the Deep Sea Floor

Early models for the process of calcium carbonate dis-solution in deep sea sediments were based on simplediagenetic models in which calcium carbonate dissolved intothe pore waters of the sediments.304 The alteration ofconcentrations from an initial saturation state of the overlyingwater to equilibrium with calcite results in diffusion of thereactants and products, causing exchange with the overlyingwaters. The process is governed by the difference betweenequilibrium concentrations and those found in the overlyingwaters, the rate of dissolution as a function of disequilibriumdiffusion rates and the thickness of the stagnant benthicboundary layer above the interface. Because of the relativelyrapid dissolution rate of calcium carbonate, the depth in thesediment calculated for equilibrium has generally been onlya few millimeters.

A major benthic process that had only casually beenconsidered for its potential influence on carbonate accumula-tion in deep sea sediments is the oxidation of organic matter.Although this reaction was widely recognized as beingimportant in coastal and continental slope sediments, which

372 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 32: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

are relatively rich in organic matter, it was generally ignoredin deep sea sediments, which usually have less than 0.2 wt% organic C. With the advent of sediment traps, however,it became apparent that significant amounts of organic matterwere reaching the sediment-water interface. Consequently,near-interfacial oxidation of organic matter is potentially amajor diagenetic process even in deep sea sediments. Adetailed review of this process has recently been made293

for the reader interested in pursing the topic further.Understanding of the factors that control the diagenesis

of deep sea carbonates that leads to their partial to completedissolution can be considered largely one of scale. Fordifferences among major ocean basins, it is clear that theprimary variable is the saturation state of the water at a givendepth, which is largely controlled by its pCO2. A secondaryinfluence is the deposition rate of siliciclastic sediments.Within ocean basins, variations in near-surface primaryproductivity and the ratio of the rain rate of organic carbonto calcium carbonate to sediments can exert a stronginfluence and, at least in some areas, lead to substantialdissolution above the saturation depth of the overlying water.This process will be explored further in the subsequentdiscussions of carbonate diagenesis in shoal-water carbonate-rich and shallow-to-intermediate depth siliciclastic sedimentswhere metabolizable organic matter is at higher concentra-tions and exerts even a greater influence on carbonatediagenesis.

The approach that has generally been used is to estimatecarbonate dissolution rates via diagenetic models thatincorporate organic matter oxidation by oxygen (assuminga C/O ratio) that produces undersaturated conditions.305 Arelationship is then established between the calcite dissolutionrate and pore water saturation state. The general kineticequation for calcite dissolution in seawater (eq 3.12166) hasbeen most commonly used.

The fundamental parameter in all models for calciumcarbonate dissolution in the deep sea is the saturation stateof pore waters. In order to determine the saturation state,not only the composition of the pore waters but also thesolubility of the calcium carbonate must be known underinsitu conditions. Therefore, many studies of carbonate chem-istry in deep sea sediment pore waters have focused on theapparent solubility behavior of carbonates in these sediments.Early results306,307indicated that, despite considerable vari-ability in the pore water chemistry of sediments at a numberof different sites in the Atlantic and Pacific oceans, the porewaters are generally close to calculated equilibrium withcalcite. Higher ion concentration products in pore waterswhere Mn2+ was detectable were observed306 indicatingpossible control by a carbonate phase other than calcite anda much higher ion concentration product in one sedimentcontaining aragonite was also found.307 Hales and Emerson308

have observed different apparent solubilities for sedimentsfrom the Ceara Rise and Ontong-Java Plateau, but the reasonsfor this remain uncertain.

Perhaps no aspect of the diagenesis of deep sea carbonateshas been more controversial than the values of rate constantand reaction order. Most investigators have modeled ratesthat demand rate constants orders of magnitude less thanthose measured for deep sea sediments in the laboratory.Also, the values for the reaction order have ranged from 1309

to 4.5.310 Even within a given study, as with apparentsolubility products, different values for the rate constant havebeen required.308,310 A puzzling observation was made by

Jahnke and Jahnke311 who found that in sediments abovethe saturation depth containing high concentrations ofcalcium carbonate, the ratio of the calcium carbonatedissolution rate to the organic matter remineralization ratewas substantially less than at other types of sites. Theysuggested that this may be the result of exchange oncarbonate particle surfaces coupled with particle mixing, butthis process has yet to be clearly substantiated. Althoughconsiderable speculation has been put forth to attempt toexplain these large variations in deep sea carbonate dissolu-tion kinetics, at this time it appears our ability to makepredictions is still quite limited and that a number ofparameters may contribute to these variations.

5.2. Dissolution of Carbonate Minerals inCarbonate-Rich Shallow Water Sediments

Many studies of the impact of chemical diagenesis on thecarbonate chemistry of anoxic sediments have focusedprimarily on the fact that sulfate reduction results in theproduction of alkalinity, which can cause precipitation ofcarbonate minerals (see previous discussion). However,during the early stages of sulfate reduction (∼2% to 35%)this reaction may not cause precipitation, but rather resultsin undersaturation with respect to carbonate minerals becausethe impact of lowered pH is greater than that of increasedalkalinity. Carbonate ion activity decreases rapidly as it is“titrated” by CO2 from organic matter decomposition leadingto a decrease in pore water saturation state. This process isevident in data for Fe-poor shallow water carbonate sedi-ments from the Bahamas251 and has been confirmed in otherstudies of Florida Bay,312-314 Checker Reef, Oahu,315 andBermuda sediments.316 There are three primary processes thathave been identified that can cause undersaturation, inaddition to the undersaturation that may result during theearly stages of sulfate reduction. These are early postdeathmicroenvironments within organisms, oxidation of organicmatter by processes preceding sulfate reduction, and oxida-tion of sulfides. These processes commonly will be mostimportant near the sediment-water interface.

Prior to the onset of sulfate reduction, extensive organicmatter degradation can occur by bacterially mediated oxygenreduction as described for deep sea sediments. The influenceof benthic bacterial activity under aerobic conditions oncarbonate mineral dissolution has been nicely demonstratedfor pore waters from sediments of the Gulf of Calvi inCorsica.272 Other reactions of probable less importance thanthose above leading to undersaturated conditions with respectto calcium carbonate near the sediment-water interfaceinclude nitrate reduction and fermentation.317

Lynn Walter and her associates312-314 demonstrated themajor quantitative importance of dissolution of calciumcarbonate from shallow carbonate sediments. Up to about50% carbonate dissolution can be driven by the sulfatereduction-sulfide oxidation process. In calcium carbonate-rich sediments, there is often a lack of reactive iron toproduce iron sulfide minerals. The sulfide that is producedby sulfate reduction then basically can only be buried indissolved form in pore waters, be oxidized, or diffuse out ofthe sediments. For most carbonate-rich sediments, theoxidative process strongly dominates the fate of the sulfide.Figure 16 shows the strong relationship that generally occurs,in carbonate muds from Florida Bay, among total carbondioxide, excess dissolved calcium (calcium at a concentrationabove that predicted from salinity), and the amount of sulfate

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 373

Page 33: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

that has been reduced. It is noteworthy that the burrowedbanks show much more extensive increases in calcium thanthe other mud banks. This is in good agreement with theobservations in Long Island Sound siliciclastic sediments,where increased bioturbation leads to increased sulfideoxidation and carbonate dissolution.318

6. Response of Carbonate-Rich Sediments to theAcidification of the Ocean Due to RisingAtmospheric pCO 2

6.1. General Principles and ConsiderationsThere is now overwhelming scientific evidence that human

activities, such as burning of fossil fuels and land usepractices, have caused a major disruption of the natural globalcarbon cycle resulting in a major increase in atmosphericpCO2.319 This has changed the global ocean from a net sourceto a sink for CO2 and a consequence of this has been anuptake by the ocean of roughly half of the excess CO2

released.2,228,320Although there remains considerable scien-tific debate regarding the extent and distribution of climatechange caused by rising atmospheric CO2, there is noquestion that the uptake of CO2 by the ocean is causing pHto decrease321 and along with it the saturation state ofseawater with respect to carbonate minerals.322 It is hypoth-esized that this decrease in saturation state is in turn leadingto a decrease in calcification by organisms and an increasein carbonate mineral dissolution.323-325 These effects arelikely to lead to major changes in the biology and chemistryof the ocean over a rather short period of time.

In considering the reactions involved between carbonateminerals and the carbonic acid-pCO2 system, there are fewimportant fundamental relationships that will be representedhere in a simple form to illustrate the basic concepts generallyfollowing the conventions of ref 320. Intuitively, it mightbe expected that the precipitation of calcium carbonate woulddecrease solution pCO2 and dissolution of calcium carbonatewould increase pCO2 because total dissolved inorganiccarbon (DIC) concentrations and total alkalinity (TA) changein this manner. However, the opposite is true as shown ineq 6.1. Consequently, the addition of CO2 to the ocean willbe buffered by calcium carbonate dissolution and an associ-ated increase in DIC and TA.

Previously the relationship of organic matter oxidation tocarbonate diagenesis was discussed. Organic matter produc-tion is also important and results in CO2 uptake fromseawater. Production and respiration can be written simplyas opposing reactions as in eq 6.2. Organic matter productioncan occur along with calcification for formation of biogeniccarbonate minerals from important organisms such as coc-colithophorids and various coralline algae. These processescan be combined resulting in eq 6.3.

Although increasing the pCO2 of the surface ocean maylead to a decrease in CaCO3 production326 and the releaseof CO2 associated with this process, this will be counterbal-anced by the formation of organic matter, which will bestimulated by higher pCO2 values and possibly increasednutrient input to the ocean as well.320 In fact, fertilization ofthe ocean is being considered as a possible way of removingCO2. The subsequent oxidation of a major portion of theproduced organic matter in the water column and at theseafloor in sediments will locally release CO2, loweringsaturation state and causing increased carbonate mineraldissolution. Thus, in the future the ocean will likely have adecrease in biogenic formation of carbonate minerals andan increase in the fraction of carbonate minerals producedthat undergo dissolution.297

6.2. The Major Ocean BasinsIt is estimated that probably 40-80% of the CaCO3 formed

in the pelagic environment dissolves in the upper watercolumn (see section 5.1.2). The calcium carbonate that doessurvive and becomes buried on the seafloor is mostly calcite.We can therefore divide the impact of lowering the saturationstate of seawater due to increasing pCO2 into basically upperocean processes, lowering the saturation state in the deepwater column, and impacts on processes occurring near thesediment-water interface of deep sea sediments above thecurrent CCD.

The upper water column, via air-sea exchange andrelatively rapid mixing, will be first impacted by rising pCO2

resulting in a probable lowering of CaCO3 production andas previously described an increase in organic matterproduction, but nutrient availability may limit changes inprimary productivity. This will result in changing the rain-rate ratio of CaCO3 C to organic C into deep waters andsediments. Oxidation of the increased organic matter willresult a rise of pCO2 in the deep ocean prior to the penetrationof CO2-rich upper waters into the deep sea. This will lowersaturation states with respect to carbonate minerals and leadto increased dissolution both in the water column andsediments. Subsequently, in a complex pattern reflecting deepwater formation, mixing, and circulation, anthropogenic CO2

will be transported from surface waters into the main volumeof seawater that resides in the deep ocean basins. This mixingand equilibration will take thousands of years.327 In conjunc-tion with the associated major decrease in deep sea sedimentcalcium carbonate accumulation, this will be the primary

Figure 16. Dissolved sulfate versus excess Ca2+ in the pore watersof carbonate-rich Florida Bay sediment: closed squares, unburrowedmud bank; open circles, burrowed mud bank. Modified from ref313, Figure 2 with permission of the authorand the Royal Society,Copyright 1993.

Ca2+ + 2HCO3- y\z

ppt

dissolCaCO3 + H2O + CO2

(6.1)

CO2 + H2O y\zproduction

respirationCH2O + O2 (6.2)

Ca2+ + 2HCO3- h CaCO3 + CH2O + O2 (6.3)

374 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 34: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

process that ultimately reduces atmospheric pCO2 followingthe current pulse of anthropogenic input.228

6.3. Shallow Water Carbonate-Rich SedimentsAs previously discussed, the shallow ocean will be most

immediately impacted by rising atmospheric pCO2 resultingin a lowering of pH and decreasing saturation state withrespect to carbonate minerals. An impact that is alreadyoccurring is a decrease in calcification rates of organismssuch as aragonitic corals. Attempts have been made toestimate the change that is expected in net shallow water(coastal margin) ecosystem calcifications rates that indicatea rather slow drop with time up until about the year 2000 to2050 followed by a rapid decrease in the net calcificationrate.328

In order to make predictions about the dissolution ofshallow water marine carbonates in the future ocean inresponse to rising atmospheric pCO2, the anticipated calciumcarbonate ion concentration product can be computed as afunction of time for surface ocean water and compared withthe stoichiometric solubility product of different calciumcarbonate minerals. An example of such calculations is givenin Figure 17.329 As discussed previously, a major difficultyoccurs with the high magnesian calcites where the influencesof magnesium and other factors lead to large certainties intheir solubility. This is of particular concern in makingpredictions about the response of shallow marine highmagnesian calcites to declining seawater pH because theseminerals are the most soluble common carbonate mineralsand therefore would be the “first responders” to a decliningsaturation state.

A simple schematic model utilizing this concept ispresented in Figure 18. In this model,329 reservoirs of varioussizes for the different magnesian calcite, aragonite, and calcitecompositions are allowed to react until they are exhaustedin sequence of their solubility. This results in the changeswith time in solution chemistry and reservoir sizes presentedin Figure 19. Two scenarios were used based on the Plummerand Mackenzie17 and “best fit” biogenic magnesian calcitestoichiometric solubilities (see Figure 2). The results of the

two scenarios are quite different reflecting the large uncer-tainties in the solubilities of high magnesian calcites, butboth do point to the large changes in dissolution expectedfor shallow water carbonates in the relatively near future dueto changing atmospheric pCO2.

7. Summary and Thoughts on Possible FutureResearch Directions

7.1. SummaryOur objective in writing this paper was to review the major

aspects of the chemistry of carbonate mineral formation anddissolution in the marine environment. In doing so, we didnot attempt to cover all that is known about carbonate mineralaqueous solution chemistry but rather focused primarily onthose aspects of direct relevance to their behavior in seawater.Sections 2 and 3 of the paper dealt with basic propertiesand results of laboratory experiments, whereas sections 4and 5 cover observations concerned with carbonate mineralformation and dissolution in the ocean. Section 6 brieflycovers the response of the ocean and, in particular, carbonateminerals to rising atmospheric carbon dioxide concentrations.

While we have attempted to be reasonably comprehensiveon the topics we addressed, there are many related subjectsthat were only touched on or not addressed to which differentauthors might have given greater attention. These include,for example, the oceanic carbon cycle, biomineralization,incorporation of trace components and isotopes used largelyas proxies for paleooceanographic conditions, studies of earlydiagenesis, and how the ocean’s chemistry and carbonateminerals have varied over geologic time. There is anextensive body of literature on all these topics and the readerwho has gained interest in the chemistry of marine carbonatesfrom this paper is encouraged to pursue these related areas.Also, it should be noted that a review paper on the carbonicacid system in seawater by Millero2 appears in this issue,and therefore, we did not cover this topic in any detail.

Laboratory research related to carbonate mineral behaviorhas primarily been aimed at determining the solubility andreaction kinetics of calcite (including high-Mg calcites) andto a lesser extent aragonite in seawater under varyingconditions. Research on dolomite formation under oceanicconditions has been much more limited due to its very slowreaction kinetics under close to STP conditions. Untilrelatively recently, most of the research focused largely onthe composition of the solution in contact with these minerals.A major challenge has been the use of equilibrium thermo-dynamic concepts as the basis for interpreting results, eventhough in most cases the system is at best metastable. Inmany instances, it is not even possible for the carbonatemineral being studied to be in metastable equilibrium withthe seawater solution because the reaction is not reversible.

Another difficulty is that most of the carbonate mineralsformed in the ocean are derived from biogenic sources wherethe composition of solution from which they are precipitatedis controlled by biologic processes often referred to in generalas “vital” effects. This can lead to biogenic carbonates thatare heterogeneous, partially hydrated, and disordered. Aprime example given for these differences between theproperties of abiotic and biotic carbonates was that of Mgcalcites. Additionally, the morphology of the biogeniccarbonates is often complex and microporous in natureleading to a disconnect between true (e.g., BET) surface areasand “reactive” surface areas. In some instances, these

Figure 17. Change in the calcium carbonate ion concentrationproduct in surface seawater at constant alkalinity as a function ofpCO2. Horizontal lines are for various solubility products. Greenis for 15 mol % Mg calcite based on various solubility estimates.The gray line is based on observations from the Great BahamaBank. Adapted from ref 329, Figure 11, Copyright 2006 withpermission from Elsevier.

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 375

Page 35: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

morphological properties can appear to override chemicalproperties such as solubility.

Considerable effort (and speculation) has gone into tryingto establish the mechanisms responsible for observed kineticbehavior between carbonate minerals and a bulk seawatersolution. It is now reasonably well-established that underconditions likely to be encountered in the ocean thesereactions are surface-controlled rather than transport-controlled. Consequently, there has been an emphasis ontrying to establish the chemical speciation of carbonatemineral surfaces under varying conditions and on identifyingsurficial mechanisms that are ultimately rate-controlling. Theadvent and increasing resolution and versatility of a varietyof analytical techniques capable of determining composition,morphology, and structure at scales of a few nanometers hasopened major opportunities for moving much of the researchin this area from speculation to direct observation.

One of the more perplexing aspects of understanding thebehavior of marine carbonates is the role of organic matter.Organic matter is ubiquitous in the marine environment. Mostsurfaces are covered with a film of adsorbed organic matterand many biogenic carbonate minerals have occluded organicmatter. The use of the term organic matter is certainly notprecise and reflects that much of this material is the resultof the breakdown to various degrees of complex once-livingtissue, of which only a small portion can be identified interms of specific organic compounds. Consequently, the inter-action and influences of “model” compounds have often beenused to try to assess the potential impact of organic matteron marine carbonates. The interactions of different organiccompounds and their influences on calcite and aragonitereaction kinetics vary widely. Observations of carbonatemineral diagenesis in marine sediments indicates that bio-logically derived organic matter may have little influenceon calcium carbonate dissolution but may play a major rolein inhibiting precipitation on sedimentary carbonates.

Near-surface seawater is supersaturated with respect tocalcite (including abiotic high-Mg calcite), aragonite, andideal dolomite. However, these minerals fail to precipitatedirectly from seawater under near-normal oceanic conditions.

In the modern ocean, the dominant source of carbonateminerals comes from a wide variety of organisms that yield“skeletal” carbonates. In the pelagic realm of the open ocean,the primary carbonate mineral formed is low-Mg calcite (<4mol % MgCO3) coming from foraminifera and coccolitho-phores, with smaller amounts of aragonite also being derivedfrom organisms such as pteropods and heteropods. Deepocean sediments contain high-Mg calcite (>4 mol %MgCO3) from benthic organisms and occasional abioticmarine cements usually as a very minor component of totalsedimentary carbonate minerals. The sources and mineralogyof shallow water carbonates are much more highly variedand complex than deep water carbonates, with aragonitebeing the must abundant mineral followed by high and lowmagnesian calcites. In some regions, such as the GreatBahama Bank, the primary source of calcium carbonate islikely the abiotic precipitation of fine-grained aragoniteneedle “muds”. Dolomite is rarely found and then usuallyonly in “special” environments that typically show elevatedsalinity, temperature, or saturation state.

The production of calcium carbonate is greater in thepelagic environment than the shallow water environment,but the extent of dissolution is substantially less in theshallow water environment. In both environments, diagenesisin sediments, where carbon dioxide is added from reminer-alization of organic matter, can play an important role incalcium carbonate dissolution. However, deep ocean watersbecome undersaturated and the general distribution ofcalcium carbonate in deep sea sediments can be related tothe variability in different ocean basins of the degree ofundersaturation at equivalent depths. The processes control-ling the finer scale distribution of deep sea carbonates includedissolution kinetics, organic matter content, and accumulationrate. Major hurdles remain to be overcome in determiningthe source of variations of apparent solubilities in pore water,and in resolving the differences in the dissolution kineticsdetermined in laboratory experiments with those obtainedfrom modeling deep sea sediment pore water profiles.

The ongoing increase in anthropogenic CO2 is having animpact on the degree of saturation of the surface ocean,

Figure 18. Model schematic of simple dissolution model representing the global coastal ocean. In the current simulations, the system wasassumed to be closed and did not exchange water with the open ocean, the atmosphere, or input via rivers and runoffs. Arrows denotefluxes. The masses of carbonate phases are shown at the bottom of each carbonate mineral reservoir in 1012 mol. The major forcing of themodel was pCO2, which increased by 2µatm year-1. Reprinted from ref 329, Figure 4, Copyright 2006 with permission from Elsevier.Also see ref 330.

376 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 36: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

generally referred to as the “acidification” of the ocean. Themost immediate impact is on the shallow water environment,which is in close contact with the atmosphere. Here rates ofcalcification may be negatively impacted, and eventually itis likely that these waters will become undersaturated withrespect to biogenic high-Mg calcites and in some regionseven aragonite. The reduction in calcification rate and theincreased dissolution of some shallow water sedimentarycarbonates will lead to an increase in the oceanic uptake ofCO2. Over longer time spans this “excess” CO2 will moreextensively impact the deep sea leading to more extensivedissolution there, and features such as the calcium carbonatecompensation depth will become less deep.

7.2. Thoughts on Possible Future ResearchDirections

Although research on the formation and dissolution ofcarbonate minerals in the ocean has been going on since atleast the late nineteenth century, there remain many basic

unanswered fundamental questions. Certainly three of themost important of these have to do with building bridgesbetween different areas of research that have substantiallyevolved separately and where attempts to relate findings haveoften met with limited success at best.

The first area where better conceptual bridging is neededis between classic bulk solution-mineral work, in which theinteraction between the solid and dissolved phases is oftenaproached using concepts derived from equilibrium thermo-dynamics, and more recent nanoscale mechanistic observa-tions. Many of these concepts were the major focus of thispaper. Future work will certainly have to focus on furtherdevelopment of experimental and analytical techniques. Inparticular, the improvement of time and spatial resolutionwill be of importance. The development of “super-resolution”methods is one possible pathway, and also the likely abilityto combine techniques in one instrument, for example, thecombination of confocal or fluorescence microscopy withinterferometry or even AFM with VSI techniques, will further

Figure 19. Model calculated responses of carbonate minerals to increasing pCO2 (see text for details). The figure illustrates changes insaturation state with respect to calcite, aragonite, and 11-18 mol % Mg calcite and the total masses of Mg calcite minerals present in thesediments throughout the simulation in scenario 1 (A-C) and scenario 2 (D-E). The same saturation state data are shown in panels A andB and panels D and E but on different scales in order to highlight the sequential dissolution of the most soluble carbonate phase presentin the sediments as well as the establishment of a metastable equilibrium between the seawater and this phase. Reprinted from ref 329,Figure 5, Copyright 2006 with permission from Elsevier. Also see ref 330.

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 377

Page 37: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

enhance our ability observe surface processes. However,being able to integrate these observations may also requirea revised theoretical framework that provides a means ofresolving relationships between surface composition and thedistribution of surface energy. Development of this theoreti-cal underpinning will also require an understanding of howoverall reactivity varies as a function of length and timescales. More powerful Monte Carlo based computer simula-tions that treat crystal dissolution and growth as many-bodyproblems also show promise. Computer simulations of thissize and quality are the prerequisite for a direct comparisonof high-resolution laboratory observations at the mineralsurface.

Such a combined approach will allow the investigation ofcrystal dissolution and growth kinetics from the quantummechanical to the macroscopic scale. Of course, modelingholds the most promise for those processes that are beyondthe reach of laboratory experiments. Where possible, modelpredictions based on molecular simulations should be testedexperimentally to provide a robust link between results fromthe nano-, micro-, and macroscales.

The second area where better conceptual bridging isneeded is between results obtained from abiotic experimentsand the characteristics of calcium carbonate “biominerals”.Clearly a major first step is a better understanding of howthe biogenic carbonates that dominate the marine environ-ment differ from experimentally studied abiotic phases. Thisdoes not mean that we have to know the details of howdifferent organisms make their carbonate minerals, but wemust develop an understanding of how the particular proper-ties of biogenic carbonates influence solubility and kineticbehavior. This understanding must accommodate the fact thatmany biogenic carbonate minerals are likely to be hetero-geneous on many different scales. As we have mentionedpreviously, this is a complex field in which even basicquestions, such as how sample preparation may influenceresults, are still outstanding.

The third area where better conceptual bridging is neededis between what goes on under relatively simple and wellcontrolled experimental conditions and what goes on in the“real” ocean. As discussed in this paper, there are often largedifferences between what is predicted and what is observed,often only indirectly. The problems in building bridges inthis area have commonly been associated with two quitedifferent communities of investigators, who have not onlydifferent experimental and observational methods but alsooften quite different philosophies about the potential of basicchemical approaches to ever resolve the behavior of verycomplex biogeochemical systems. However, until suchapproaches can be better brought to bear on these problems,our potential for a better understanding of the processesgoverning what we observe will be necessarily limited.

8. AcknowledgementsPreparation of this paper was in part supported by a grant

from the Chemical Sciences, Geosciences and BiosciencesDivision, Office of Basic Energy Sciences, Office of Science,U.S. Department of Energy (Grants DE-FG02-00ER15033and DE-FG02-06ER15816; J.W.M.) and the Louis andElizabeth Scherck Endowed Chair in Oceanography (J.W.M.).J.W.M. thanks Hanna Broussard for her assistance inpreparing the final draft of this paper. R.S.A. and A.L. wouldlike to thank Michael Vinson, Li Zhang, and Kevin Davisof Rice University for numerous discussions and insights

involving carbonate reaction kinetics and for supplyinggenerous assistance with figures as well. R.S.A. and A.L.also gratefully acknowledge funding provided by the Officeof Naval Research (ONR), the Department of Energy (DOENo. DE FG07-01ER63295), the Department of DefenseMultidisciplinary University Research Initiative (MURI)program, and the Nanoscale Science and Engineering Initia-tive of the National Science Foundation (No. EEC-0118007).We especially thank Fred T. Mackenzie for his manyinsightful comments in his review of this paper. Furtherimprovements were made based on comments of an anony-mous reviewer.

References(1) Morse, J. W.; Mackenzie, F. T.Geochemistry of Sedimentary

Carbonates; Elsevier: Amsterdam, 1990.(2) Millero, F. J.Chem. ReV. 2007, 107, 308-341.(3) Shannon, R.Acta Crystallogr., Sect. A1976, 32, 751.(4) Sverjensky, D. A.Geochim. Cosmochim. Acta1984, 48, 1127.(5) Doerner, H. A.; Hoskins, W. M.J. Am. Chem. Soc.1925, 47, 662.(6) Morse, J. W.; Bender, M. L.Chem. Geol.1990, 82, 265.(7) Morse, J. W; Mucci, A.; Millero, F. J.Geochim. Cosmochim. Acta

1980, 44, 85.(8) Mucci, A. Am. J. Sci.1983, 238, 780.(9) Computed from SUPCRT92: Johnson, J. W.; Oelkers, E. H.;

Helgeson, H. C.Comput. Geosci.1992, 18, 899.(10) Mucci, A. Aquat. Geochem.2004, 10, 139.(11) Spek, A. L.J. Appl. Crystallogr.2003, 36, 7.(12) de Villiers, J. P. R.Am. Mineral.1971, 56, 758.(13) Reeder, R. J.ReV. Mineral. 1983, 11, 1.(14) Graf, D. L.Am. Mineral.1961, 46, 1283.(15) Klein, C.The Manual of Mineral Science, 22nd ed.; John Wiley &

Sons: New York, 2002.(16) Mackenzie, F. T.; Bischoff, W. D.; Bishop, F. C.; Loijens, M.;

Schoonmaker, J.; Wollast, R.ReV. Mineral. 1983, 11, 97.(17) Plummer, L. N.; Mackenzie, F. T.Am. J. Sci.1974, 274, 61.(18) Walter, L. M.; Morse, J. W.Geochim. Cosmochim. Acta1984, 48,

1059.(19) Thorstenson, D. C.; Plummer, L. N.Am. J. Sci.1977, 277, 1203.(20) Berner, R. A.Am. J. Sci.1978, 278, 1475.(21) Garrels, R. M.; Wollast, R.Am. J. Sci.1978, 278, 1469.(22) Lafon, G. M.Am. J. Sci.1978, 278, 1455.(23) Thorstenson, D. C.; Plummer, L. N.Am. J. Sci.1978, 278, 1478.(24) Mucci, A.; Morse, J. W.Geochim. Cosmochim. Acta1983, 47, 217.(25) Mucci, A.; Morse, J. W.Geochim. Cosmochim. Acta1984, 48, 815.(26) Mucci, A.; Morse, J. W.; Kaminsky, M. S.Am. J. Sci.1985, 285,

289.(27) Bischoff, W. D.; Mackenzie, F. T.; Bishop, F. C.Geochim.

Cosmochim. Acta1987, 51, 1413.(28) Bischoff, W. D.; Bishop, F. C.; Mackenzie, F. T.Am. Mineral.1983,

68, 1183.(29) Busenberg, E.; Plummer, L. N.Geochim. Cosmochim. Acta1989,

53, 1189.(30) Lafon, G. M. InFluid-Mineral Interactions: A Tribute to H. P.

Eugster; Spencer, R. J., Chou, I-M., Eds.; Special Publication SeriesNo. 2; The Geochemical Society: San Antonio, TX, 1990; p 23.

(31) Bertram, M. A.; Mackenzie, F. T.; Bishop, F. C.; Bischoff, W. D.Am. Mineral.1991, 76, 1889.

(32) Bischoff, W. D.; Sharma, S. K.; Mackenzie, F. T.Am. Mineral.1985,71, 581.

(33) Bischoff, W. D.; Bertram, M. A.; Mackenzie, F. T.; Bishop, F. C.Carbonates EVaporites1993, 8, 82.

(34) Vasconcelos, C.; McKenzie, J. A.; Bernasconi, S.; Grujic, D.; Tien,A. J. Nature1995, 377, 220.

(35) Roberts, J. A.; Bennett, P. C.; Gonza´lez, L. A.; Macpherson, G. L.;Milliken, K. L. Geology2004, 32, 277.

(36) Goldsmith, J. R.; Heard, H. C.J. Geol.1961, 69, 45.(37) Graf, D. L.; Goldsmith, J. R.J. Geol.1956, 64, 173.(38) Sperber, C. M.; Wilkinson, B. H.; Peacor, D. R.J. Geol.1984, 92,

609.(39) Mucci, A. Can. Mineral.1991, 29, 113.(40) Garrels, R. M.; Thompson, M. E.; Siever, R.Am. J. Sci.1960, 258,

402.(41) Navrotsky, A.; Capobianco, C.Am. Mineral.1987, 72, 782.(42) Chai, L.; Navrotsky, A.; Reeder, R. J.Geochim. Cosmochim. Acta

1995, 59, 939.(43) Plummer, L. N.; Busenberg, E.Geochim. Cosmochim. Acta1982,

46, 1011.

378 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 38: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

(44) Sass, E.; Morse, J. W.; Millero, F. J.Am. J. Sci.1983, 283, 218.(45) Berner, R. A.Am. J. Sci.1976, 276, 713.(46) Tribble, J. S.; Mackenzie, F. T.Aquat. Geochem.1998, 4, 337.(47) Lasaga, A. C.Kinetic Theory in the Earth Sciences; Princeton

University Press: Princeton, NJ, 1998.(48) Sjoberg, L.Geochim. Cosmochim. Acta1976, 40, 441.(49) Morse, J. W.Am. J. Sci.1978, 278, 344.(50) Morse, J. W.; Arvidson, R. S.Earth Sci. ReV. 2002, 58, 51.(51) Reddy, M. M.; Nancollas, G. H.J. Colloid Interface Sci.1971, 36,

166.(52) Nancollas, G. H.; Reddy, M. M.J. Colloid Interface Sci.1971, 37,

824.(53) Nagy, K. L.; Blum, A. E; Lasaga, A. C.Am. J. Sci.1991, 291, 649.(54) Nielsen, A. E.Kinetics of Precipitation; Pergamon: Oxford, U.K.,

1964.(55) Blum, A. E.; Yund, R. A.; Lasaga, A. C.Geochim. Cosmochim. Acta

1990, 54, 283.(56) Schott, J.; Brantley, S.; Crerar, D.; Guy, C.; Borcsik, M.; Willaime,

C. Geochim. Cosmochim. Acta1989, 53, 373.(57) Lasaga, A. C.ReV. Mineral. 1995, 8, 23.(58) Burton, W. K.; Cabrera, N.; Frank, F. C.Philos. Trans. R. Soc.

London, Ser. A1951, 243, 299.(59) Shiraki, R.; Rock, P. A.; Casey, W. H.Aquat. Geochem.2000, 6,

87.(60) Higgins, S. R.; Boram, L. H; Eggleston, C. M.; Coles, B. A.;

Compton, R. G.; Knauss, K. G.J. Phys. Chem. B2002, 106, 6696.(61) Arvidson, R. S.; Collier, M.; Davis, K. J.; Vinson, M. D.; Amonette,

J. E.; Luttge, A.Geochim. Cosmochim. Acta2006, 70, 583.(62) Van Cappellen, P.; Charlet, L.; Stumm, W.; Wersin, P.Geochim.

Cosmochim. Acta1993, 57, 3505.(63) Arakaki, T.; Mucci, A.Aquat. Geochem.1995, 1, 105.(64) Pokrovsky, O. S.; Schott, J.EnViron. Sci. Technol.2002, 36, 426.(65) Duckworth, O. W.; Martin, S. T.Geochim. Cosmochim. Acta2003,

67, 1787.(66) Binnig, G.; Quate, C. F.; Gerber, C.Phys. ReV. Lett.1986, 56, 930.(67) Marti, O.; Drake, B.; Hansma, P. K.Appl. Phys. Lett.1987, 51, 484.(68) Hillner, P. E.; Gratz, A. J.; Manne, S.; Hasma, P. K.Geology1992,

20, 359.(69) Gratz, A. J.; Hillner, P. E.; Hasma, P. K.Geochim. Cosmochim. Acta

1993, 57, 491.(70) Dove, P. M.; Platt, F. M.Chem. Geol.1996, 127, 331.(71) Liang, Y.; Baer, D. R.Surf. Sci.1997, 373, 275.(72) Davis, K. J.; Dove, P. M.; De Yoreo, J. J.Science2000, 290, 1134.(73) Vinson, M. D.; Luttge, A.Am. J. Sci.2006, 305, 119.(74) Higgins, S. R.; Jordan, G.; Eggleston, C. M.Geochim. Cosmochim.

Acta 2002, 66, 3201.(75) Hu, X.; Grossie, D. A.; Higgins, S. R.Am. Mineral.2005, 90, 963.(76) Weyl, P. K.J. Geol.1958, 66, 163.(77) Terjesen, S. G.; Erga, O.; Thorsen. G.; Ve, A.Chem. Eng. Sci.1961,

74, 277.(78) Meyer, H. J.J. Cryst. Growth1984, 66, 639.(79) Cabrera, N.; Vermilyea, D. A.Growth and Perfection of Crystals;

Chapman & Hall: London, 1958.(80) Gutjahr, A.; Dabringhaus, H.; Lacmann, R.J. Cryst. Growth1996,

158, 310.(81) Luttge, A. In Nanoscale Structure and Assembly at Solid-Fluid

Interfaces. Vol. I. Interfacial StructuresVersus Dynamics; Liu, X.-Y., De Yoreo, J. J., Eds.; Kluwer Academic Publishers: Boston, MA,2004; p 209.

(82) White, A. F., Brantley, S. L., Eds.Chemical Weathering Rates ofSilicate Minerals; Reviews in Mineralogy 31; Mineralogical Societyof America: Washington, DC, 1995.

(83) Brady, P. V.; Walther, J. V.Geochim. Cosmochim. Acta1989, 53,2823.

(84) Stumm, W.; Wieland, E. InAquatic Chemical Kinetics; Stumm, W.,Ed.; Wiley-Interscience: New York, 1990.

(85) Stumm, W.; Wollast, R.ReV. Geophys.1990, 28, 53.(86) Gautier, J.-M.; Oelkers, E. H.; Schott, J.Geochim. Cosmochim. Acta

1994, 58, 4549.(87) Oelkers, E. H.; Schott, J.; Devidal, J. L.Geochim. Cosmochim. Acta

1994, 58, 2011.(88) Lasaga, A. C. InChemical Weathering Rates of Silicate Minerals;

White, A. F., Brantley, S. L., Eds; Reviews in Mineralogy 31;Mineralogical Society of America: Washington, DC, 1995; p 23.

(89) Casey, W. H.; Ludwig, C. InChemical Weathering Rates of SilicateMinerals; White, A. F., Brantley, S. L., Eds; Reviews in Mineralogy31; Mineralogical Society of America: Washington, DC, 1995; p87.

(90) Dove, P. M. InChemical Weathering Rates of Silicate Minerals;White, A. F., Brantley, S. L., Eds; Reviews in Mineralogy 31;Mineralogical Society of America: Washington, DC, 1995; p 87.

(91) Blum, A. E.; Stillings, L. L. InChemical Weathering Rates of SilicateMinerals; White, A. F., Brantley, S. L., Eds; Reviews in Mineralogy31; Mineralogical Society of America: Washington, DC, 1995; p235.

(92) Oelkers, E. H.Geochim. Cosmochim. Acta2002, 65, 3703.(93) Blum, A. E.; Lasaga, A. C. InAquatic Surface Chemistry: Chemical

Processes at the Particle-Water Interface; Stumm, W., Ed.; Wiley:New York, 1987; p 255.

(94) Hochella, M. F., Jr.; Banfield, J. F. InChemical Weathering Ratesof Silicate Minerals; White, A. F., Brantley, S. L., Eds; Reviews inMineralogy 31; Mineralogical Society of America: Washington, DC,1995; p 353.

(95) Brantley, S. L.; White, A. F.; Hodson, M. E. InGrowth, Dissolutionand Pattern Formation in Geosystems; Jamtveit B., Meakin P., Eds;Kluwer: Amsterdam, 1999; p 291.

(96) Brantley, S. L.; Mellott, N. P.Am. Mineral.2000, 85, 1767.(97) Jeschke, A. A.; Dreybrodt, W.Geochim. Cosmochim. Acta2002,

66, 3055.(98) Luttge, A. Am. Mineral.2005, 90, 1776.(99) Gautier, J.-M.; Oelkers, E. H.; Schott, J.Geochim. Cosmochim. Acta

2001, 65, 1059.(100) Brunauer, S.; Emmet, P. H.; Teller, E.J. Am. Chem. Soc.1938, 60,

309.(101) Anbeek, C.Geochim. Cosmochim. Acta1992, 56, 1461.(102) Hochella, M. F., Jr. InMineral Surfaces; Vaughan, D. J., Pattrick,

R. A. D., Eds.; Mineral Society Series 5; Chapman & Hall: London,1995; p 17.

(103) Arvidson, R. S.; Ertan, I. E.; Amonette, J. E.; Lu¨ttge, A. Geochim.Cosmochim. Acta2003, 67, 1623.

(104) Lichtner, P. C., Steefel, C. I., Oelkers, E. H., Eds.ReactiVe Transportin Porous Media; Reviews in Mineralogy 34; Mineralogical Societyof America: Washington, DC, 1997; p 407.

(105) Luttge, A.; Winkler, U.; Lasaga, A. C.Geochim. Cosmochim. Acta2003, 67, 1099.

(106) Beig, M. S.; Lu¨ttge, A.Geochim. Cosmochim. Acta2006, 70, 1402.(107) Dove, P. M.; Han, N.; De Yoreo, J. J.Proc. Natl. Acad. Sci. U. S. A.

2005, 102, 15357.(108) Hellmann, R.; Tisserand, D.Geochim. Cosmochim. Acta2006, 70,

364.(109) Luttge, A. J. Electron Spectrosc. Relat. Phenom.2005, 150, 288.(110) Velbel, M. A.Chem. Geol.1989, 78, 245.(111) Velbel, M. A.Water Resour. Res.1990, 26, 3049.(112) White, A. F. InChemical Weathering Rates of Silicate Minerals;

White, A. F., Brantley, S. L., Eds; Reviews in Mineralogy 31;Mineralogical Society of America: Washington, DC, 1995.

(113) Badiozamani, K.; Mackenzie, F. T.; Thorstenson, D. C.J. Sediment.Petrol. 1977, 47, 529.

(114) Kazmierczak, T. F.; Tomson, M.; Nancollas, G. H.J. Phys. Chem.1982, 86, 103.

(115) Zhong, S.; Mucci, A.Chem. Geol.1989, 78, 283.(116) Zuddas, P.; Mucci, A.Geochim. Cosmochim. Acta1998, 62, 757.(117) Glover, E. D.; Sippel, R. F.Geochim. Cosmochim. Acta1967, 31,

603.(118) Kinsman, J. J.; Holland, H. D.Geochim. Cosmochim. Acta1969,

33, 1.(119) Katz, A.Geochim. Cosmochim. Acta1973, 37, 1563.(120) Fuchtbauer, H.; Hardie, L. A.Annu. Meet. Geol. Soc. Am.1976, 8,

877.(121) Mucci, A.Geochim. Cosmochim. Acta1987, 51, 1977.(122) Bischoff, W. D.Aquat. Geochem.1998, 4, 321.(123) Gaffey, S.J. Sediment. Petrol.1988, 58, 397.(124) Moberly, R.Sedimentology1968, 11, 61.(125) Lippmann, F.Sedimentary Carbonate Minerals; Springer-Verlag:

New York, 1973.(126) Mucci, A.Geochim. Cosmochim. Acta1986, 50, 2255.(127) Arvidson, R. S.; Mackenzie, F. T.Aquat. Geochem.2000, 6, 249.(128) Mucci, A.; Morse, J. W.Am. J. Sci.1985, 285, 306.(129) Busenberg, E.; Plummer, L. N.Geochim. Cosmochim. Acta1985,

49, 713.(130) Brady, P. V.; Krumhansl, J. L.; Papenguth, H. W.Geochim.

Cosmochim. Acta1996, 60, 727.(131) Lorens, R. B.Geochim. Cosmochim. Acta1981, 45, 553.(132) Pingitore, N. E.; Eastman, M. P.; Sandidige, K. O.; Freiha, B.Mar.

Chem.1988, 25, 107.(133) McBride, M. B.J. Soil Sci. Soc. Am.1979, 43, 693.(134) Franklin, M. L.; Morse, J. W.Mar. Chem.1983, 12, 241.(135) Mucci, A.Geochim. Cosmochim. Acta1988, 52, 1859.(136) Plummer, L. N.; Busenberg, E.Geochim. Cosmochim. Acta1987,

54, 1393.(137) Stumm, W.; Leckie, J. O.AdV. Water Pollut. Res.1970, 2, 26.(138) Leckie, J. O.; Stumm, W. InWater Quality ImproVement by Physical

and Chemical Processes; Gloyna, E. F., Eckenfelder, W. W., Jr.,Eds.; University of Texas Press: Austin, TX, 1970; p 237.

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 379

Page 39: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

(139) Martens, C. S.; Harriss, R.Geochim. Cosmochim. Acta1970, 34,621.

(140) de Kanel, J.; Morse, J. W.Geochim. Cosmochim. Acta1978, 42,1335.

(141) Suess, E.Geochim. Cosmochim. Acta1970, 34, 157.(142) Berner, R. A.; Westrich, J. T.; Graber, R.; Smith, J.; Martens, C. S.

Am. J. Sci.1978, 278, 816.(143) Zullig, J. J.; Morse, J. W.Geochim. Cosmochim. Acta1988, 52, 1667.(144) Braitsch, O.Salt Deposits. Their Origin and Composition; Springer-

Verlag: New York, 1971.(145) Sonnenfeld, P.Brines and EVaporites; Academic Press: Orlando,

FL, 1964.(146) Cloud, P. E., Jr.U.S. Geol. SurV. Prof. Pap.1962, 350.(147) Pytkowicz, R. M.Am. J. Sci.1973, 273, 515.(148) He, S.; Morse, J. W.Geochim. Cosmochim. Acta1993, 57, 3533.(149) Morse, J. W.; He, S.Mar. Chem.1993, 41, 291.(150) Hardie, A. L.Geology1996, 24, 279.(151) Stanley, S. M.; Hardie, L. A.Palaeogeogr., Palaeoclimatol., Palaeo-

ecol.1998, 144, 3.(152) Ries, J. B.Geology2004, 32, 981.(153) Morse, J. W.; Wang, Q.; Tsio, M. Y.Geology1997, 25, 85.(154) Zhang, J.; Nancollas, G. H.J. Cryst. Growth1990, 106, 181.(155) Plummer, L. N.; Wigley, T. M. L.; Parkhurst, D. L.Am. J. Sci.1978,

278, 179.(156) Sjoberg, E. L.; Rickard, D. T.Chem. Geol.1984, 42, 119.(157) Rickard, D. T.; Sjo¨berg, E. L.Am. J. Sci.1983, 283, 815.(158) Sjoberg, E. L.; Rickard, D. T.Geochim. Cosmochim. Acta1983, 47,

2281.(159) Plummer, L. N.; Parkhurst, D. L.; Wigley, T. M. L. InChemical

ModelingsSpeciation, Sorption, Solubility and Kinetics in AqueousSystems; Jenne, E., Ed.; American Chemical Society: Washington,DC, 1979; p 537.

(160) Sjoberg, E. L.Stockholm Contrib. Geol.1978, 32, 1.(161) Gutjahr, A.; Dabringhaus, H.; Lacmann, R.J. Cryst. Growth1996,

158, 296.(162) Sjoberg, E. L.; Rickard, D. T.Geochim. Cosmochim. Acta1984, 48,

485.(163) Peterson, M. N. A.Science1966, 154, 1542.(164) Berger, W. H.Science1967, 156, 383.(165) Milliman, J. D. In The Fate of Fossil Fuel CO2 in the Oceans;

Andersen, N. R., Malahoff, A., Eds.; Plenum Press: New York, 1977;p 641.

(166) Morse, J. W.; Berner, R. A.Am. J. Sci.1972, 272, 840.(167) Berner, R. A.; Morse, J. W.Am. J. Sci.1974, 274, 108.(168) Kier, R. S.Geochim. Cosmochim. Acta1980, 44, 241.(169) Morse, J. W.; de Kanel, J.; Harris, J.Am. J. Sci.1979, 279, 482.(170) Busenberg, E.; Plummer, L. N. InStudies in Diagenesis; Mumton,

F. A., Ed.; U.S. Geological Survey Bulletin 1578: Washington D.C.,1986; p 139.

(171) Land, L. S.J. Sediment. Petrol.1967, 37, 914.(172) Wollast, R.; Reinhard-Derie; D. InThe Fate of Fossil Fuel CO2 in

the Oceans; Andersen, N.R., Malahoff, A., Eds.; Plenum Press: NewYork, 1977; p 479.

(173) Walter, L. M.; Morse, J. W.Geochim. Cosmochim. Acta1985, 49,1503.

(174) Banfield, J. F.; Hamers, R. J. InGeomicrobiology; Banfield J. F.,Nealson, K. H., Eds.; Reviews in Mineralogy and Geochemistry 35;Mineralogical Society of America: Washington, DC, 1997; p 81.

(175) Berner, R. A.; Lasaga, A. C.; Garrels, R. M.Am. J. Sci.1983, 283,641.

(176) Garrels, R. M.; Lerman, A.Am. J. Sci.1984, 284, 989.(177) Westbroek, P.Life as A Geological Force; Norton & Co: New York,

1991.(178) Lowenstam, H. A.; Weiner, S.Biomineralization; Oxford University

Press: New York, 1989.(179) Dove, P. M., De Yoreo, J. J., Weiner, S., Eds.Biomineralization;

Reviews in Mineralogy and Geochemistry 54; Mineralogical Societyof America: Washington, DC, 2003.

(180) Boquet, E.; Boronat, A.; Ramos-Cormenzana, A.Nature1973, 246,527.

(181) Novitsky, J. A.Geomicrobiology1981, 2, 375.(182) Pentecost, A.; Bauld, J.Geomicrobiology1988, 6, 129.(183) Banfield, J. F., Nealson, K. H., Eds.Geomicrobiology: Interactions

Between Microbes and Minerals; Reviews in Mineralogy, Vol. 35;Mineralogical Society of America: Washington, DC, 1997.

(184) Lower, S. K.Am. J. Sci.2005, 305, 752.(185) Escher, A.; Characklis, W. G. InBiofilms; Characklis W. G., Marshall

K. C., Eds.; J. Wiley & Sons: New York, 1990; p 445.(186) Lower, S. K.; Hochella, M. F., Jr.; Beveridge, T. J.Science2001,

292, 1360.(187) Luttge, A.; Conrad, P. G.Appl. EnViron. Microbiol. 2004, 70, 1627.(188) Cabrera, N.; Levine, M. M.Philos. Mag.1956, 1, 450.(189) Cabrera, N.; Levine, M. M.; Plaskett, J. S.Phys. ReV. 1954, 96, 1153.

(190) Lasaga, A. C.; Lu¨ttge, A. Science2001, 291, 2400.(191) Lasaga, A. C.; Lu¨ttge, A. Eur. J. Mineral.2003, 15, 603.(192) Jordan, G.; Rammensee, W.Geochim. Cosmochim. Acta1998, 62,

941.(193) Little, B. J.; Wagner, P.; Heart, K.; Ray, R.; Lavoie, D.; Nealson, K.

H.; Aguilar, C. Corrosion 97; National Association of CorrosionEngineers: Houston, TX, 1997; p 215.

(194) Little, B. J.; Wagner, P.; Heart, K.; Ray, R.; Lavoie, D.; Nealson, K.H.; Aguilar, C.Biodegradation1998, 9, 1.

(195) Luttge, A.; Zhang, L.; Nealson, K. H.Am. J. Sci.2005, 305, 766.(196) Davis, K. J.; Lu¨ttge, A. Am. J. Sci.2005, 305, 727.(197) Davis, K. J.; Lu¨ttge, A.Geobiology, submitted for publication, 2007.(198) Conrad, P. G.; Lu¨ttge, A. Unpublished results.(199) Banfield, J. F., Navrotsky, A., Eds.Nanoparticles and the EnViron-

ment; Reviews in Mineralogy and Geochemistry, Vol. 44; Mineral-ogical Society of America: Washington, DC, 2001.

(200) Banfield, J. F.; Zhang, H. InNanoparticles and the EnVironment;Banfield, J. F., Navrotsky, A., Eds.; Reviews in Mineralogy andGeochemistry, Vol. 44; Mineralogical Society of America: Wash-ington, DC, 2001; p 1.

(201) Navrotksy, A. InNanoparticles and the EnVironment; Banfield, J.F., Navrotsky, A., Eds.; Reviews in Mineralogy and Geochemistry,Vol. 44; Mineralogical Society of America: Washington, DC, 2001;p 73.

(202) Kitano, Y.; Hood, D. W.J. Oceanogr. Soc. Jpn.1962, 3, 141.(203) Kitano, Y.; Hood, D. W.Geochim. Cosmochim. Acta1965, 29, 29.(204) Chave, K. E.Science1965, 148, 1723.(205) Chave, K. E.; Suess, E.Trans. N. Y. Acad. Sci.1967, 29, 991.(206) Suess, E. Calcium Carbonate Interaction with Organic Compounds.

Ph.D. Dissertation, Lehigh University, Bethlehem, PA, 1968.(207) Suess, E.Geochim. Cosmochim. Acta1973, 37, 2435.(208) Wollast, R. InCarbonate Cements; Bricker, A. P., Ed.; Johns Hopkins

University Press: Baltimore, MD, 1971; p 264.(209) Berner, R. A.Principles of Chemical Sedimentology; McGraw-Hill:

New York, 1971.(210) Pytkowicz, R. M.J. Geol.1965, 73, 196.(211) Berner, R. A.Geochim. Cosmochim. Acta1975, 39, 489.(212) Mitterer, R. M.Science1968, 162, 1498.(213) Mitterer, R. M. InCarbonate Cements; Bricker O. P., Ed.; Johns

Hopkins University Press: Baltimore, MD, 1971; p 95.(214) Mitterer, R. M.Geochim. Cosmochim. Acta1972, 36, 1407.(215) Trichet, J.Compt. Rend.1968, 1492.(216) Sjoberg, E. L.Contrib. Geol.1978, 332, 1.(217) Hare, P. E.Geol. Soc. Am. Spec. Pap.1961, 68.(218) Hare, P. E.Science1963, 139, 216.(219) Simkiss, K.Some Proc. Malac. Soc. London1960, 34, 89.(220) Mitterer, R. M.Bull. Mar. Sci.1978, 28, 173.(221) Carter, P. W.Geochim. Cosmochim. Acta1978, 42, 1239.(222) Carter, P. W.; Mitterer, R. M.Geochim. Cosmochim. Acta1978, 42,

1231.(223) Mitterer, R. M.; Cunningham, R., Jr. InCarbonate Cements;

Schneidermann, N., Harris, P. M. Eds.; Special Publication (Societyof Economic Paleontologists and Mineralogists) 36; Society ofEconomic Paleontologists and Mineralogists: Tulsa, OK, 1985, p17.

(224) Jackson, T. A.; Bischoff, J. L.J. Geol.1971, 79, 493.(225) Muller, P. J.; Suess, E.Geochim. Cosmochim. Acta1977, 41, 941.(226) Lahann, R. W.; Campbell, R. C.Geochim. Cosmochim. Acta1978,

44, 629.(227) Zullig, J. J. Interaction of Organic Compounds with Carbonate

Mineral Surfaces in Seawater and Related Solutions. Ph.D. Disserta-tion, Texas A&M University, College Station, TX, 1985.

(228) Feely, R. A.; Sabine, C. L.; Lee, K.; Berelson, W.; Kleypass, J.;Fabry, V. J.; Millero, F. J.Science2004, 305, 362.

(229) Andersson, A. J.; Mackenzie, F. T.; Lerman, A.Global Biogeochem.Cycles2006, 20, GB1S92.

(230) Sabine, C. L.; Heimann, M.; Artaxo, P.; Bakker, D.; Chen, C.-T. A.;Field, C. B.; Gruber, C.; LeQue´re, C.; Prinn, R. G.; Richey, J. E.;Lankao, P. R.; Sathaye, J.; Valentini, R. InSCOPE 62, The GlobalCarbon Cycle: Integrating Humans, Climate, and the Natural World;Field, H. B., Rapauch, M. R., Eds.; Island Press: Washington, DC,2004.

(231) Berner, R. A.; Honjo, S.Science1981, 211, 940.(232) Lerman A. L.; Dacey M. F. InSuspended Solids in Water; Gibbs, R.

A., Ed.; Plenum Press: New York, 1974; p 17.(233) Lerman, A. L. Geochemical Processes: Water and Sediment

EnVironment; Wiley-Interscience: New York, 1979.(234) Honjo, S. InDissolution of Deep-Sea Carbonates; Sliter, W., Be,

A., Berger, W. H., Eds.; Cushman Foundation for ForaminiferalResearch, Special Publication No. 13; Cushman Foundation forForaminiferal Research: New York, 1975; p 114.

(235) Honjo, S.Marine Micropaleontology1976, 1, 65.

380 Chemical Reviews, 2007, Vol. 107, No. 2 Morse et al.

Page 40: Calcium Carbonate Formation and Dissolutionepsc511.wustl.edu/Morse2007.pdfCalcium Carbonate Formation and Dissolution John W. Morse,* Rolf S. Arvidson, and Andreas Lu¨ttge Department

(236) Milliman, J. D.Recent Marine Carbonates 1, Marine Carbonates;Springer-Verlag: Berlin, 1974.

(237) Milliman, J. D.; Muller, J. Sedimentology1973, 20, 29.(238) Milliman, J. D.; Muller, J. In The Fate of Fossil Fuel CO2 in the

Oceans; Anderson, N. R., Malahof, A., Eds.; Plenum Press: NewYork, 1977.

(239) Garrels, R. M.; Wollast, R.Am. J. Sci.1978, 278, 1469.(240) Milliman, J. D.; Troy, P. J.; Balch, W. M.; Adams, A. K.; Li, Y.-H.;

Mackenzie, F. T.Deep Sea Res.1999, 46, 1653.(241) Droxler, A. W.; Morse, J. W.; Kornicker, W. A.J. Sediment. Petrol.

1988, 58, 120.(242) Droxler, A. W.; Morse, J. W.; Glaser, K. S.; Haddad, G. A.; Baker,

P. A. Mar. Geol.1992, 100, 277.(243) Reid, R. P.; Macintyre, I. G.; Post, J. E.J. Sediment. Petrol.1992,

62, 145.(244) Andrews, J. E.; Christidis, S.; Dennis, P. F.J. Sediment. Res.1997,

67, 531.(245) Bathurst, R. G. C.Carbonate Sediments and Their Diagenesis;

Developments in Sedimentology 12; Elsevier: Amsterdam, 1975.(246) Scoffin, T. P.An Introduction to Carbonate Sediments and Rocks;

Blackie: Glasgow, 1987.(247) Milliman, J. D.; Freile, D.; Steinen, R. P.; Wilber, R. J.J. Sediment.

Petrol. 1993, 63, 589.(248) Lowenstam, H. A.; Epstein, S.J. Geol.1957, 65, 364.(249) Loreau, J. P.Sediments Aragonitiques et leur Genese.Memoires du

Museum National d’Histoire Naturelle, Naturelle se´rie. Serie C,Geologie; Editions du Muse´um: Paris, Tome XLVII, 1982.

(250) Morse, J. W.; Gledhill, D. K.; Millero, F. J.Geochim. Cosmochim.Acta 2003, 67, 2819.

(251) Morse, J. W.; Zullig, J. J.; Bernstein, L. D.; Millero, F. J.; Milne, P.;Mucci, A.; Choppin, G. R.Am. J. Sci.1985, 285, 147.

(252) Debenay, J. P.; Andre, J. P.; Lesourd, M.Mar. Geol. 1999, 157,159.

(253) Land, L. S.; Behrens, E. W.; Frishman, S. A.J. Sediment. Petrol.1979, 49, 1269.

(254) Fabricius, F. H.Origin of Marine Ooids. Contributions to Sedimen-tology; E. Schweizerbartche Verlagsbuchhandlung: Stuttgart, Ger-many, 1977.

(255) Sorby, H. C.Proc. Geol. Soc. London1879, 35, 56.(256) Mitterer, R. M.Science 1968, 167, 1499.(257) Suess, E.; Fu¨tterer, D.Sedimentology1972, 19, 129.(258) Bathurst, R. G. C.Annu. ReV. Earth Planet. Sci.1974, 2, 257.(259) Bathurst, R. G. C.Mar. Geol.1967, 5, 89.(260) Weyl, P. K.Int. Conf. Trop. Oceanogr.1967, 178.(261) Davies, P. J.; Bubela, B.; Furguson, J.Sedimentology1978, 25, 73.(262) Ferguson, J.; Bubela, B.; Davies, P. J.Chem. Geol.1978, 22, 285.(263) Bricker, O. P.Carbonate Cements; Johns Hopkins University Studies

in Geology 19; Johns Hopkins Press: Baltimore, MD, 1971.(264) Schneidermann, N., Harris, P. M., Eds.Carbonate Cements; Society

of Economic Paleontologists and Mineralogists: Tulsa, OK, 1985.(265) Harris, P. J. InCarbonate Cements;Schneidermann, N., Harris, P.

M., Eds.; Society of Economic Paleontologists and Mineralogists:Tulsa, OK, 1985; p 79.

(266) Morse, J. W.; Mucci, A.Sediment. Geol.1984, 40, 287.(267) Ginsburg, R. N.; Schroeder, J. H.; Shinn, E. A. InCarbonate Cements;

Bricker, O. P., Ed.; Johns Hopkins University Studies in Geology19; Johns Hopkins Press: Baltimore, MD, 1971; p 54.

(268) Bathurst, R. G. C.Sedimentology1987, 34, 749.(269) Glover, E. D.; Pray, L. C. InCarbonate Cements; Bricker, O. P.,

Ed.; Johns Hopkins University Studies in Geology 19; Johns HopkinsPress: Baltimore, MD, 1971; p 80.

(270) Given, R. K.; Wilkinson, B. H.J. Sediment. Petrol.1985, 55, 109.(271) Mullins, H. T.; Wise, S. H., Jr.; Gardulski, A. F.; Hinchey, E. J.;

Masters, P. M.; Siegel, D. I.Sedimentology1985, 32, 473.(272) Moulin, E.; Jordens, A.; Wollast, R.Proc. Prog. Belgium Oceanogr.

Res.1985, 196.(273) Alexandersson, E. T.; Milliman, J. D.J. Sediment. Petrol.1981, 51,

1309.(274) Thorstenson, D. C.; Mackenzie, F. T.Geochim. Cosmochim. Acta.

1974, 38, 1.(275) Gaillard, J.-F.; Jeandel, C.; Michard, G.; Nicolas, E.; Renard, D.Mar.

Chem.1986, 18, 233.(276) Ritger, S.; Carson, B.; Suess, E.Geol. Soc. Am. Bull.1987, 98, 147.(277) Chafetz, H. S.J. Sediment. Petrol.1986, 56, 812.(278) Morrow, D. W.Geosci. Can.1982, 9, 5.(279) Goldsmith, J. R.J. Geol.1953, 61, 439.(280) Land, L. S.Aquat. Geochem.1998, 4, 361.(281) Wanless, H. R.J. Sediment. Petrol.1979, 49, 437.(282) Gunatilaka, A.; Saleh, A.; Al-Temeemi, A. Presented at the SEPM

Mid-Year Meeting, Golden, Colorado, 1985, p 38.(283) Baker, P. A.; Burns, S. J.AAPG Bull.1985, 69, 1917.(284) Freidman, G. M.Soc. Econ. Paleontol. Mineral. Spec. Publ.1980,

28, 69.

(285) Hardie, L. A. J. Sediment. Petrol. 1987, 57, 166.

(286) Compton, J. S.Geology1988, 16, 318.(287) Mazzullo, S. J.J. Sediment. Res.2000, 70, 10.(288) Arvidson, R. S.; Mackenzie, F. T.Aquat. Geochem.1997, 2, 273.(289) Arvidson, R. S.; Mackenzie, F. T.Am. J. Sci.1999, 299, 257.(290) Vasconcelos, C.; McKenzie, J. A.J. Sediment. Res.1997, 67, 378.(291) Morse, J. W. InChemistry of Marine Water and Sediments;

Gianguzza, A., Pelizzetti, E., Sammartano, S., Eds.; Springer: NewYork, 2002; p 165.

(292) Morse, J. W. InTreatise on Geochemistry, V7: Sediments, Diagenesisand Sedimentary Rocks; Mackenzie, F. T., Ed.; Elsevier: Amsterdam,2003; p 67.

(293) Ben-Yaakov, S.Limnol. Oceanogr.1973, 18, 86.(294) Rigter, S.; Calrson, B.; Suess, E.Geol. Soc. Am. Bull.1987, 98, 147.(295) Roberts, H. H.; Aharon, P.Geo-Mar. Lett.1994, 14, 135.(296) Naehr, T. H.; Rodriguez, N. M.; Bohrmann, G.; Paull, C. K.; Botz,

R. Proc. Ocean Drill. Prog.2000, 164, 285.(297) Archer, D.; Maier-Reimer, E.Nature1994, 367, 260.(298) Li, Y.-H.; Takahashi, T.; Broecker, W. S.J. Geophys. Res.1969,

74, 5507.(299) Morse, J. W.; Berner, R. A.Am. J. Sci.1972, 274, 638.(300) Mekik, F. A.; Loubere, P. W.; Archer, D. E.Global Biogeochem.

Cycles2002, 16, 1.(301) Thunell, R. C.; Honjo, S.Mar. Geol.1981, 40, 237.(302) Troy, P. J.; Li, Y.-H.; Mackenzie, F. T.Aquat. Geochem.1997, 3, 1.(303) Betzer, P. R.; Byrne, R. H.; Acker, J. G.; Lewis, C. S.; Jolley, R. R.;

Feely, R. A.Science1984, 226, 1074.(304) Takahashi, T.; Broecker, W. S. InThe Fate of Fossil Fuel CO2 in

the Oceans; Anderson, N. R., Malahof, A., Eds.; Plenum Press: NewYork, 1977; p 455.

(305) Archer, D. E.; Morford, J. L.; Emerson, S. R.Global Biogeochem.Cycles2002, 16, BG1288.

(306) Emerson, S.; Jahnke, R.; Bender, M.; Froelich, P.; Klinkhammer,G.; Bowser, C.; Setlock, G.Earth Planet. Sci. Lett.1980, 49, 57.

(307) Sayles, F. L.Mar. Chem.1980, 9, 223.(308) Hales, B.; Emerson, S.Geochim. Cosmochim. Acta1997, 61, 501.(309) Hales, B.; Emerson, S.Earth Planet. Sci. Lett.1997, 148, 317.(310) Wenzho¨fer, F.; Holby, O.; Kohls, O.Deep-Sea Res.2001, 48, 1741.(311) Jahnke, R. A.; Jahnke, D. B.Geochim. Cosmochim. Acta2004, 68,

47.(312) Walter, L. M.; Burton, E. A.Am. J. Sci.1990, 290, 601.(313) Walter, L. M.; Bischof, S. A; Patterson, W. P.; Lyons, T. W.Philos.

Trans. R. Soc. London, Ser. A1993, 344, 27.(314) Ku, T. C. W.; Walter, L. M.; Coleman, M. L.; Blake, R. E.; Martini,

A. M. Geochim. Cosmochim. Acta1999, 63, 2529.(315) Tribble, G. W. Early Diagenesis in a Coral Reef Framework. Ph.D.

Dissertation, University of Hawaii, Honolulu, 1990.(316) Wollast, R.; Mackenzie, F. T. Unpublished data.(317) Aller, R. C.AdV. Geophys.1980, 22, 237.(318) Aller, R. C.; Rude, P. D.Geochim. Cosmochim. Acta1988, 52, 751.(319) Barnola, J. M.; Aklin, M.; Procheron, J.; Reynaud, D.; Schwander,

D.; Stauffer, B.Tellus, Ser. B1995, 47, 264.(320) Lerman, A.; Mackenzie, F. T.Aquat. Geochem.2005, 11, 345.(321) Caldiera, K.; Wickett, M. E.Nature2003, 425, 365.(322) Andersson, A. J.; Mackenzie, F. T.; Lerman, A.Am. J. Sci.2005,

305, 875.(323) Gao, K.; Aruga, Y.; Asada, K.; Ishihara, T.; Akano, T.; Kiyohara,

M. Mar. Biol. 1993, 117, 129.(324) Gattuso, J.-P.; Frankignoulle, M.; Bourge, I.; Romaine, S.; Budde-

meier, R. W.Global Plant Change1998, 18, 37.(325) Doney, D. C.Sci. Am.2006, 294, 58.(326) Orr, J. C.; Fabry, V. J.; Aumont, O.; Bopp, L.; Doney, S. C.; Feely,

R. A.; Gnanadesikan, A.; Gruber, N.; Ishida, A.; Joos, F.; Key, R.M.; Lindsay, K.; Maier-Reimer, E.; Matear, R.; Monfray, P.;Mouchat, A.; Najjar, R. G.; Plattner, G.-K.; Rodgers, K. B.; Sabine,C. L.; Sarmiento, J. L.; Schlitzer, R.; Slater, R. D.; Totterdell, I. J.;Weirig, M.-F.; Yamanaka, Y.; Yool, A.Nature2005, 437, 681.

(327) Archer, D.; Kheshgi, H.; Maier-Reimer, E.Global Biogeochem.Cycles1998, 12, 259.

(328) Mackenzie, F. T.; Andersson, A.; Lerman, A.; Ver, L. M. InTheSea Vol. 13: The Global Coastal Ocean; Robinson, A. R., Brink,K. H., Eds.; Harvard University Press: Cambridge, MA, 2004; p133.

(329) Morse, J. W.; Andersson, A. J.; Mackenzie, F. T.Geochim.Cosmochim. Acta2006, 70, 5814.

(330) Andersson, A. J. Climate Change and Anthropogenic Effects onShallow-water Carbonate Biogeochemistry. M.S. Thesis, Universityof Hawaii, Honolulu, 2003.

CR050358J

Calcium Carbonate Formation and Dissolution Chemical Reviews, 2007, Vol. 107, No. 2 381