1 s2.0 S0012825201000836 Main
1 s2.0 S0012825201000836 Main
www.elsevier.com/locate/earscirev
Abstract
Among the most important set of chemical reactions occurring under near Earth surface conditions are those involved in the
dissolution of sedimentary carbonate minerals. These minerals comprise about 20% of Phanerozoic sedimentary rocks. Calcite
and, to a significantly lesser extent, dolomite are the major carbonate minerals in sedimentary rocks. In modern sediments,
aragonite and high-magnesian calcites dominate in shallow water environments. However, calcite is by far the most abundant
carbonate mineral in deep sea sediments. An understanding of the factors that control their dissolution rates is important for
modeling of geochemical cycles and the impact of fossil fuel CO2 on climate, diagenesis of sediments and sedimentary rocks. It
also has practical application for areas such as the behavior of carbonates in petroleum and natural gas reservoirs, and the
preservation of buildings and monuments constructed from limestone and marble. In this paper, we summarize important
findings from the hundreds of papers constituting the large literature on this topic that has steadily evolved over the last half
century. Our primary focus is the chemical kinetics controlling the rates of reaction between sedimentary carbonate minerals and
solutions. We will not attempt to address the many applications of these results to such topics as mass transport of carbonate
components in the subsurface or the accumulation of calcium carbonate in deep sea sediments. Such complex topics are clearly
worthy of review papers on their own merits. Calcite has been by far the most studied mineral over a wide range of conditions and
solution compositions. In recent years, there has been a substantial shift in emphasis from measuring changes in solution
composition, to determine ‘‘batch’’ reaction rates, to the direct observation of processes occurring on mineral surfaces using
techniques such as atomic force microscopy (AFM). However, there remain major challenges in integrating these two very
different approaches. A general theory of surface dissolution mechanisms, currently lacking (although see Lasaga and Luttge
[Science 291 (2001) 2400]), is required to satisfactorily relate observations of mineral surfaces and the concentration of dissolved
components. Studies of aragonite, high-magnesian calcites, magnesite, and dolomite dissolution kinetics are much more limited
in number and scope than those for calcite, and provide, at best, a rather rudimentary understanding of how these minerals are
likely to behave in natural systems. Although the influences of a limited number of reaction inhibitors have been studied,
probably the greatest weakness in application of experimental results to natural systems is understanding the often profound
influences of ‘‘foreign’’ ions and organic matter on the near-equilibrium dissolution kinetics of carbonate minerals. D 2002
Elsevier Science B.V. All rights reserved.
*
Corresponding author. Tel.: +1-409-845-9630; fax: +1-409-845-9631.
E-mail address: [email protected] (J.W. Morse).
0012-8252/02/$ - see front matter D 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 0 1 2 - 8 2 5 2 ( 0 1 ) 0 0 0 8 3 - 6
52 J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84
1. Introduction and background would render, in some cases, the paper virtually unread-
able. Citations will be made using examples as to what
1.1. Overview and general considerations are, in our opinion, the most important, clearly stated,
and representative papers on the topic.
One of the more basic pieces of knowledge, as
anyone who takes an introductory geology course 1.2. A brief pre-1970 historical perspective
learns, is that if you put a drop of dilute HCl on a
rock and it fizzes, the rock is limestone. Thus, starts The importance of carbonate mineral dissolution in
the learning process about the seemingly simple sub- sediments (e.g., Murray and Renard, 1891), during
ject of carbonate mineral dissolution kinetics; a topic diagenesis and lithification (e.g., Friedman, 1964;
on which hundreds of research papers have been Gross, 1964; Land, 1967), and during the evolution
written and basic aspects of which still remain highly of sedimentary rocks (see extensive discussions in
controversial. Bathurst, 1975; Moore, 1989; Morse and Mackenzie,
Why all this interest in carbonate mineral dissolu- 1990) has long been recognized. These observations
tion kinetics? The answer to the question is simple. inevitably demanded an understanding of the factors
These phases are plentiful in sedimentary rocks and that controlled reaction rates between carbonate min-
modern sediments, and the dissolution process impacts erals and natural waters. However, it was not until
a vast range of important topics. Examples include the about the 1960s that a significant effort was begun to
fate of fossil fuel CO2, carbonate accumulation in determine experimentally the rates at which carbonate
marine sediments, global geochemical cycles, the mineral dissolution occurs as a function of solution
preservation of monuments and buildings, and petro- composition (e.g., Weyl, 1958, 1965; Terjesen et al.,
leum reservoir characteristics. In the early 1980s, 1961; Akin and Lagerwerff, 1965; Berner, 1967;
Morse (1983) wrote a review article on the more Schmalz, 1967; Nestaas and Terjesen, 1969). As these
general topic of calcium carbonate dissolution and lines of investigation began to evolve, it also became
precipitation kinetics and again, more briefly, visited apparent that major improvements were necessary in
the topic with Fred Mackenzie (Morse and Mackenzie, the ability to predict carbonate mineral solubility in
1990) in the late 1980s. For this paper, we have chosen natural waters (Garrels et al., 1961). There was con-
to cover the dissolution kinetics of primarily abiotic currently a growing appreciation of the complexities
aragonite, calcite, high magnesian calcite, magnesite, of biogenic carbonates, such as aragonite and magne-
and dolomite. These minerals comprise the vast major- sian calcite, and their behavior during dissolution
ity of carbonate minerals found in sediments and (Friedman, 1964; Land, 1966, 1967; Schroeder and
sedimentary rocks. Our primary objective is to distill Siegel, 1969), as well as the relationship of these
from a large number of diverse papers into a concise phases to sedimentary dolomite and dolomitization
summary of carbonate mineral dissolution kinetics. As reaction pathways (Kinsman, 1965; Friedman and
such, this paper is primarily limited to laboratory Sanders, 1967; Land and Epstein, 1970).
studies of mineral dissolution kinetics and does not Much attention was also focused on processes con-
cover the extensive literature dealing with dissolution trolling carbonate deposition in deep sea sediments.
in generally complex natural systems. Special attention was given as to how dissolution
In addition to our personal collections of papers on kinetics influence the calcium carbonate compensation
this topic, a search was made for papers in the Geo Ref depth (CCD, where calcium carbonate ceases to be a
and Chem Abstracts data bases, and then of the refer- major sediment component). The experiments of Peter-
ences in these papers. Although we hope this represents son (1966) and Berger (1967), in which calcite spheres
a reasonably complete literature search, there always and assemblages of foraminifera were hung at different
remains the possibility that, lurking in some private depths in the Pacific Ocean, were extremely important.
company’s confidential documents or in a government They revealed that the extent of calcite dissolution in
agency’s internal ‘‘gray literature’’ report, there may be seawater was not simply proportional to saturation state
important data that have been missed. No attempt will and, therefore, the dissolution rate was not diffusion
be made to cite all references on a given topic, as this controlled at modest degrees of undersaturation.
J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84 53
Their findings, and the large amount of subsequent Except at conditions very close to equilibrium,
supporting research discussed in this paper, have been generally either the forward or reverse rate is much
a source of considerable discomfort to those who wish greater than the opposing reaction so that the net rate
carbonate mineral dissolution rates to simply be is an excellent approximation of the dominant uni-
directly proportional to the extent of undersaturation directional rate (R). The reaction rate will increase
(i.e., kinetically first order and thus easily adapted for with increasing degree of disequilibrium until the
modeling purposes). Attempts have been made to reaction mechanism changes to transport-controlled
explain away or to simply ignore the complexities of kinetics.
near-equilibrium carbonate mineral reaction kinetics Here the saturation state will be given as X, which
(e.g., Edmond, 1971; Hales and Emerson, 1997). is the ratio of the ion activity product (IAP) to the
However, none of these arguments have been based solubility product (Ksp) for the solid (X = IAP/Ksp).
on direct experimental observations and, consequently, [Note that various representations of the saturation
must be viewed with skepticism in light of the very state are given. A common one is the ‘‘saturation
substantial body of experimental work carried out in index’’, SI, which is generally logX.] The extent of
both the basic chemical and geochemical communities. disequilibrium is then simply the difference between
X and 1. Thus, for undersaturation it must range from
0 to 1, but for supersaturation it can range from 0 up
2. Theoretical considerations to numbers substantially > 1.
The most commonly used equation in geosciences
2.1. Rate equations to describe the rate of carbonate mineral dissolution
(e.g., Morse and Berner, 1972) is
The reaction between a solid and solution is
dmcalcite A
observable by a change in the mass of solid and a R¼ ¼ k ð1 XÞn ð3Þ
change in solution composition. Because these repre- dt V
sent ‘‘net’’ changes, it is not usually possible to obtain where R is the rate in Amol m 2 h 1 and m is moles
‘‘absolute’’ rates of dissolution and precipitation, but of calcite, t is time, A is the total surface area of the
only to observe the difference between the opposing solid, V is the volume of solution, k is the rate constant
reactions. The concept of kinetic equilibrium rests on (note that A, V, and k are often combined into a single
the idea that at equilibrium the rates of the opposing constant k*) and n is a positive constant known as the
reactions are equal so that no change is observable ‘‘order’’ of the reaction. This simple empirical equa-
with time even though dissolution and precipitation tion has the advantage that in a plot of logR vs.
are continually occurring. log(1 X) the intercept will be k* and the slope n.
Lasaga (1998, pp. 82– 93) has cogently discussed
logR ¼ nlogð1 XÞ þ k* ð4Þ
this concept, the ‘‘principle of detailed balancing’’,
and here his more relevant points for calcite dissolu- Lasaga (1998) found that by detailed balancing (see
tion are summarized following his mode of notation. previous definition), he could obtain Eq. (5), which
For the simple reversible first order reaction, at was used by Sjöberg (1976), who found n = 1/2 for
equilibrium the concentrations of a reactant ([A]) calcite.
and product ([P]) are such that the equilibrium con-
dmcalcite A
stant Keq is equal to the ratio of the forward (k + ) and R¼ ¼ n
k Ksp ð1 X n Þ ð5Þ
dt V
reverse (k ) rate constants.
Eq. (5) is similar to Eq. (3). However, they are not
kþ simply related. Since the observed rate for each
AX P ð1Þ equation (R3 and R5), at a given saturation state, must
k
be equal, Eq. (6) gives the relationship between
equations. (Note that for n3 = n5 = 1, k3/k5 = Ksp.)
kþ ½Pe
¼ ¼ Keq ð2Þ
k ½Ae k3 ð1 XÞn3 ¼ k5 Ksp
n5
ð1 X n5 Þ
54 J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84
The growth rate law can be modified to apply to from which the following relationship can be derived:
dissolution as well and provides a bridge between
thermodynamics and kinetics. (The following treat- ceq c ¼ ceq ð1 expðDG=RT ÞÞ ð13Þ
ment is generally after that of Lasaga, 1998, modified
as appropriate to the purposes of this paper.) The At high undersaturations the rate becomes inde-
general form of this equation is pendent of DG and a ‘‘dissolution plateau ’’ is reached,
whereas close to equilibrium (small DG ) the approx-
Y
R ¼ k0 ðAeEa =RT Þ ani i f ðDGÞ ð7Þ imation given in Eq. (14) is valid and produces the
i ‘‘linear region’’.
where A is now reactive surface area (cf. Eq. (5)), Ea f ðDGÞ ¼ DG=RT ð14Þ
is the apparent activation energy, T is absolute tem-
perature, R is the ideal gas constant, and ai incorpo- A more general form of Eq. (13) that is often used,
rates ions that may contribute inhibitory or catalytic where n p 1, is Eq. (15) (Lasaga, 1998).
effects. It should be noted that implicit in this treat-
ment is that inhibitory and catalytic ions influence the f ðDGÞ ¼ ð1 expðnDG=RT ÞÞ ð15Þ
apparent overall rate constant but not the reaction
order. This is not always true as inhibitors such as The three major rate equations will be called here
phosphate can strongly influence reaction order in the general (Eq. (3)), detailed balancing (Eq. (5)) and
addition to the rate constant (Morse, 1973, 1974). (It thermodynamic rate equations (Eqs. (7) and (15)).
also assumes that their influence can be simply Because their forms are quite different, mathematical
described by their solution phase activity raised to relationships among them are not simple (e.g., Eq.
some power.) Since their influence often is on the (6)). In order to demonstrate how they differ func-
surface of the solid, in many cases explicit adsorption tionally, Fig. 1 has been constructed. This was done
isotherm equations would be more appropriate. f (DG) using different values of n. It should be kept in mind
is a function of the Gibbs free energy of disequili- that in Eq. (3), n is a ‘‘true’’ or ‘‘classical’’ reaction
brium ( = 0 at equilibrium). order, but that the n’s in Eqs. (5) and (15) are not. A
Let us consider Eq. (7) in a simplified form. First, range in saturation from equilibrium to 0.1 was used.
assume a simple solution where no inhibitory or In order to make results comparable, rates were
catalytic ions are present and that T is constant. Since normalized to the rate at X = 0.1. This results in a
rate ranging from 0 at equilibrium to 1 at X = 0.1; the
DG ¼ RT lnX ð8Þ equations produce a straight line for n = 1. For the
we can combine terms to let general rate equation the shape of the plot becomes
increasingly non-linear and concave with increasing n.
k V¼ k0 ðeEa =RT Þ ð9Þ At high values of n this can result in an apparent
change of solubility. For the detailed balancing equa-
and then letting k W= 2.303RT tion, values of n < 1 produce concave curves and for
R ¼ Ak Vf ðk WlogXÞ ð10Þ values of n > 1 convex curves. For the thermodynamic
rate equation, curves of increasing convexity are
For a constant reactive surface area, and if f is such produced with increasing n. It should be kept in mind
that k VVcan be combined with k Vyielding k, the that normalization to X = 0.1 rates hides differences in
following simple equation results. absolute rates for a set of values of n in a given
equation. It should also be noted that in cases where
R ¼ k f ðSIÞ ð11Þ reaction rate is linear with respect to X, all equations
J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84 55
R ¼ kð½CO2 2
3 eq ½CO3 bulk water Þ ð16Þ
from the surface. When this process is rate-limiting, represents the critical undersaturation for a change
the reaction is said to be diffusion controlled. Steps 2 – from surface to diffusion controlled reaction rates.
6 occur on the surface of the solid and when one of Much of the theoretical work on surface controlled
them is rate controlling the reaction is said to be reaction rates rests on the concept that surfaces are
surface controlled. As Berner (1978) pointed out, very energetically highly heterogeneous. Because reaction
soluble minerals tend to undergo reactions that are rates at sites of different energy on the surface can be
diffusion-controlled, whereas relatively insoluble min- expected to vary exponentially with site energy, the
erals have surface controlled reaction rates until very relative abundance of high-energy sites exerts a major
high degrees of disequilibrium are obtained. Carbonate influence on reaction rate. Two general classes of
minerals fall into the latter category. Surface controlled surface sites exist. The first is dislocations, with screw
reactions tend to be non-linear with respect to satu- dislocations being probably the most important. The
ration state while diffusion controlled reactions are second class of sites is those associated with steps,
linear with respect to saturation state. Although most ledges and ‘‘kinks’’. These lead to models of surface
literature on the topic is concerned with crystal growth reactions such as those proposed by Burton et al.
rather than dissolution, many of the same basic con- (1951) (see Lasaga, 1998, for extensive discussion;
cepts are applicable to both processes (e.g., Nielson, see also Fig. 2). These high-energy sites are also
1964; Lasaga, 1998). favorable adsorption sites for reaction inhibitors.
A solution can either be close to stationary (‘‘stag- Their interaction with these sites can cause large
nant’’) or moving (laminar or turbulent flow) with changes in reaction rates even at relatively low per-
respect to the solid surface. In moving solutions, centages of surface coverage.
advective transport is substantially faster than molec- Unfortunately there is currently no sound theoret-
ular diffusion. Therefore, the thickness of a stagnant ical framework for relating directly the form of
boundary layer, where water movement is slow rela- reaction rate – saturation state equations to surface
tive to diffusive transport, between the solid surface mechanisms. It is even more unfortunate that there
and bulk turbulent solution can control the rate of has been what at best might be considered highly
reaction. Mathematically this can be demonstrated speculative efforts to do so by some researchers
from Fick’s first law studying carbonate mineral dissolution kinetics, as
will be discussed later in this paper. Also to be
c c
eq bulk discussed later are relatively recent approaches that
J ¼ D ð17Þ
x are being used to observe directly surfaces on a near
atomic scale. These approaches show promise for ning frequency. Dove and Platt (1996) have estimated
better understanding of processes occurring on carbo- an overall range of 10 6 to 10 10 mol m 2 s 1.
nate mineral surfaces. More importantly, they point to discrepancies between
AFM rates computed from migrating step velocities
2.3. New approaches to studying surface processes and those measured by conventional powder methods,
and to long-standing problems regarding the treatment
Regardless of their working differences, the various of reactive surface area in these calculations. Liang et
rate equations described above all involve the notion al. (1996b) and Lea et al. (2001) have also measured
that the distance from equilibrium, as described by X individual step velocities of etch pits. Despite the
DGr
ð¼ e RT Þ, is a driving force in controlling the rate of the detail in these measurements, the overall relationship
overall reaction, and explicitly or implicitly emphasize among etch pit step velocities and the phenomeno-
the role of dissolved components in terms of the logical rate laws described above is unclear. These
overall rate mechanism. As will be discussed in more issues are discussed in more detail below.
detail shortly, in the study of the surface controlled A more recent development is the application of
reaction, there has been increasing interest in under- optical interferometry to the study of mineral surface
standing how the macroscopically observed rate is kinetics (MacInnis and Brantley, 1992; Lüttge et al.,
explicitly controlled by processes on the mineral sur- 1999). The interferometer is essentially an optical
face. This interest has come about to a large extent microscope equipped with a Mirau objective that
because of the advent of the atomic force microscope produces a series of interference fringes. The inter-
(AFM, Binnig et al., 1986; Marti et al., 1987). AFM action of these fringes with the mineral surface is used
permits direct, atomic scale in aquo observation of to produce a topographic map of the surface. Dis-
individual mineral faces. It can provide angstrom-level solution rates are measured through time-lapse scans
vertical resolution; lateral resolutions range from the of a reacted (oriented) mineral surface:
micron- to subnanometer-scale, but with increasing
resolution coming at the expense of a smaller field of Dh
¼ vðhklÞ ð18Þ
view. The instrument is ideally suited to resolving Dt
relatively flat surfaces whose overall relief is less than
1
0.5 Am. Since its introduction, AFM has seen broad r ¼ vðhklÞ V ð19Þ
application in the study of mineral – water interac-
tion involving quartz (Gratz et al., 1991), feldspars Here, Dh is the difference in surface height at a given
(Hochella et al., 1990; Drake and Hellmann, 1991), point on the mineral surface after elapsed time Dt,
zeolites (Weisenhorn et al., 1990), clays (Lindgreen et v(hkl ) is the (surface-normal) velocity, V is the molar
al., 1991), and oxides (Johnsson et al., 1991), and volume, and r is the computed rate. Interferometry
AFM techniques are now well established. Application simultaneously combines high vertical resolution
to carbonate minerals thus far has focused entirely on (]0.5 nm), large vertical scan range, and a large field
calcite (Hillner et al., 1992a,b; Gratz et al., 1993; Dove of view, and thus allows imaging of near-macroscopic
and Platt, 1996; Liang et al., 1996a,b; Davis et al., surface features. Rates determined by surface-normal
2000; Lea et al., 2001). retreat of relatively large areas of the mineral surface
Despite the capabilities of AFM, application to can also be compared directly (in units of moles per
mineral reaction rates must satisfy certain criteria unit area per unit time) to those of conventional
(Dove and Platt, 1996). Because of the limited dura- powder experiments.
bility of its cantilever tip, AFM cannot approach the
time scales of extended experiments conducted with 2.4. Complexities introduced by the carbonic acid
relative simplicity in batch or even flow-through system
reactors, imposing a limit with respect to the measure-
ment of extremely slow reactions. Conversely, rates The dissolution of most salts simply results in the
must also be sufficiently slow to be resolvable through release of their constituent cations and anions (e.g.,
comparison of raster images acquired at a fixed scan- NaCl ! Na + + Cl ) to solution. However, the disso-
58 J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84
lution of carbonate minerals necessarily involves the These difficulties have been the focus of much
carbonic acid system. Therefore, parameters that con- discussion and controversy over the years. In fact,
trol carbonate ion activity such as pH, PCO2, and some research groups such as Plummer et al. (e.g.,
alkalinity also must be considered. For example, the Plummer and Wigley, 1976; Plummer et al., 1978,
relationship between aCO23 and PCO2 and pH is given 1979; Busenberg and Plummer, 1986) have largely
by Eq. (20), where it can be observed that PCO2 and pH based their interpretations and models for calcite
can co-vary over a range of values and still yield the dissolution on a very non-traditional framework
same carbonate ion activity. focused on carbonic acid system parameters rather
than saturation state. These complications resulting
logaCO2
3
¼ logK** þ log PCO2 þ 2pH ð20Þ from the carbonic acid system also carry over to the
calcite surface and the relative abundances of surface-
K** is the product of the Henry’s law constant for CO2
species. This will subsequently be discussed in more
and the first and second dissociation constants of
detail.
carbonic acid (e.g., see Morse and Mackenzie, 1990).
These relationships introduce several practical and
theoretical difficulties to the study of carbonate min-
3. Calcite dissolution kinetics
eral dissolution kinetics. In the region where the
dissolution rate is controlled by diffusive transport
3.1. Dissolution kinetics in simple solutions at STP
in solution, the equilibrium concentration of ions near
the mineral surface must be determined. However,
3.1.1. General considerations
parameters such as PCO2 and pH will not be the same
Natural waters have a huge range in composition
as in the bulk solution. While it may be possible to
over which carbonate minerals will react with them. In
make a reasonably good estimate of the equilibrium
this section, we address the kinetics of calcite dis-
carbonate ion activity, in solutions where the calcium
solution in relatively dilute waters of simple compo-
ion activity is much greater than carbonate ion activity
sition. These waters can have a very wide range of pH
(Eq. (21)), the activities of other
values, PCO2s and saturation states. The major param-
Ksp eter of interest is saturation state and how it is related
aCO 2 ¼ ð21Þ to carbonic acid system species and calcium.
3 aCa2þ
Numerous studies have demonstrated that at ex-
species such as bicarbonate ions, carbonic acid and treme undersaturations diffusion controlled dissolution
CO2(aq) will be dependent on near-surface PCO2 and kinetics prevail. As equilibrium is approached, there is
pH. Each of these species will, therefore, have a a transition region to surface-controlled dissolution
different chemical potential gradient between the near kinetics, and then a region of changing surface-con-
surface region and the bulk solution. trolled reaction mechanisms until equilibrium is
Further complications (e.g., Weyl, 1958; Berner reached (Fig. 3). Older studies tended to largely focus
and Morse, 1974; Sjöberg, 1976; Sjöberg and Rick- on greater distances from equilibrium where changes
ard, 1983; Rickard and Sjöberg, 1983; Dreybrodt and in bulk solution composition could be readily moni-
Buhmann, 1991; Dreybrodt et al., 1996; Liu and tored and manipulated. Many more recent studies,
Dreybrodt, 1997; Zhang and Grattoni, 1998) include: utilizing advanced techniques such as atomic force
microscopy (AFM), have been concerned with devel-
1. pH involves the hydrogen ion whose diffusion oping a better understanding of reaction mechanisms
coefficient is about three times, faster than that on mineral surfaces.
of the other diffusing species, Although studies that measure bulk solution chem-
2. demands of electro-neutrality at any point istry can provide good information on the relation-
along the transport region, ships between solution composition and reaction
3. reactions between the diffusing species, and rates, they remain largely non-mechanistic in their
4. the fact that these reactions involve their interpretation. Surface studies provide interesting
kinetic rates. insights into the types of reactions that are occurring,
J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84 59
Fig. 3. Schematic representation of rate controlling mechanisms for calcite dissolution as a function of pH and temperature. (Based on Sjöberg
and Rickard, 1983.)
but are hard to extrapolate to net reaction rates. A The slope of the linear fit to the data ( 0.95) can be
major challenge is to find better ways to integrate assumed within the uncertainty of the data to be unity,
these approaches into meaningful and practically indicating a first order reaction with respect to aH +.
useful models. This relationship appears to hold up over about 5
orders of magnitude in rate. It is therefore possible to does the transition pH, from about 4.75 in pure KCl to
represent the rate (R) by the simple equation 6 in KCl containing 10 2 M Ca2 +. Sjöberg and
Rickard (1985) argue that because flow rates in many
R ¼ ka nHþ ð22Þ natural systems are relatively slow and Ca2 + concen-
tration are often substantial that an end-member trans-
At 25 jC, Plummer et al. (1978) gave a value for k of port-controlled dissolution rate for calcite may persist
about 50 10 3 cm s 1 and n = 1, whereas, Sjöberg over a substantial pH range.
and Rickard (1984a) obtained a value for k of about Compton et al. (1989) found that under very acidic
3 10 3 cm s 1 (for a mid-range diffusion boun- (pH < 4) conditions, dissolution control passed from
dary layer (DBL) thickness of 30 Am) and n = 0.9. At H + diffusion control to a first order heterogeneous
a pH of 4, these values produce a difference in reaction of H + at the surface. Their data could be
predicted rate of close to seven times faster for the modeled assuming a finite rate of heterogeneous
results of Plummer et al. (1978) than those of Sjöberg reaction according to Eq. (23), where the rate constant
and Rickard (1984a). However, the results are close to has a value of 0.043 F 0.015 cm s 1 and [H + ]s is the
identical if a lower limiting value of 4 Am is used for surface H + concentration.
the DBL, illustrating the importance of the difficult
task of assessing the thickness of the DBL. R ðmol cm2 s1 Þ ¼ k½Hþ s ð23Þ
It is interesting that the rate in this region is
independent of PCO2. A possible explanation for this Yet, another complication that has received con-
is that as pH decreases, PCO2 must increase (Eq. (20)). siderable attention is that of reaction kinetics among
Over the diffusion-controlled range of dissolution, the components of the carbonic acid system. The
PCO2 must be quite high, eventually exceeding 1 atm hydration/dehydration kinetics of dissolved CO2 are
at very low pH, as evidenced by the ‘‘fizzing’’ of sufficiently slow that under some circumstances they
calcite in acidic solutions. For example, using Eq. (20) may play an important role in dissolution reactions.
and setting PCO2 = 1 atm and pH = 4, results in a This appears most likely to occur in systems with
carbonate ion activity of 10 10.85. This would require relatively low V/A ratios such as subsurface waters
a ridiculous equilibrium calcium ion activity of close (e.g., see extensive discussions in Dreybrodt and
to 50 for a purely transport-controlled system! Buhmann, 1991; Dreybrodt et al., 1996; Liu and
Since this very simple calculation leads to unsat- Dreybrodt, 1997).
isfactory implications, a better explanation must be For small (]10 Am) particles, it has been found
found. Sjöberg and Rickard (1984a), using the rotat- that the stagnant boundary layer can be assumed
ing disk method and mixed kinetic theory, were able infinitely thick and that hydrodynamic conditions
to show that the transition to H + dependence occurs are not of great importance (e.g., Nielson, 1964). In
when H + penetrates the DBL and first reaches the the case of larger particles and large surfaces, this is
surface. Reactions in the surface boundary layer result not true and hydrodynamic conditions can become a
in a surface H + concentration value several orders of factor in controlling the rate of dissolution (see dis-
magnitude less than that in solution. This provides an cussion in Morse, 1983). For biogenic carbonates with
avenue for avoiding extremely high near-surface PCO2 complex microporous structures, the relationships
and calcium ion activity values at equilibrium, which between ‘‘geometric’’ surface area based on particle
is obviously not the case. size and ‘‘reactive’’ surface area are complex (Walter
Sjöberg and Rickard (1985) investigated the issue and Morse, 1984b).
of the influence of dissolved calcium on dissolution
kinetics in the H + -dependent region in an elegant 3.1.3. The transition region to surface controlled
paper. They demonstrated that in this region solution dissolution
calcium concentrations did not influence dissolution Following the general conceptual classification
rates. The concentration of calcium did, however, scheme of Van Name and Hill (1916), Sjöberg and
influence the pH at which the transition to H + depend- Rickard divided their studies of carbonate reaction
ence occurred. As calcium concentration increases so kinetics into the three general categories (see Fig. 3)
J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84 61
3.1.4. Near-equilibrium surface controlled dissolution Helgeson et al. (1984), they pointed out that the basic
Rickard and Sjöberg (1983) made the important rate equation for dissolution under substantially
observation that chemically controlled rate constants undersaturated conditions
(kC) depend differently on surface areas for different Y n
calcites (e.g., Iceland spar and precipitate powders). R ¼ k Se ai i ð29Þ
They argue that this is due to differences in the types i
and abundances of reactive surface sites on the differ- could be modified (following Blum and Lasaga, 1987)
ing calcites. Compton et al. (1986) also concluded that to include the influence of dislocations by setting
changes in surface morphology (roughness), arising (
X DG þ
from differences in sample preparation techniques or kSe ¼ m exp
p
the dissolution reaction itself, could cause differences perfect surface sites
kT
in measured dissolution rate. The distressing result of
X )
this is that there can be no general equation that is DG þd
þ exp þ ð30Þ
applicable to all calcites that simply relates surface dislocation surface sites
kT
area and solution composition! It also implies that the
influences of inhibitors will differ for different cal- where m is a frequency factor and k is the Boltzman
cites. The detailed study of the surfaces of dissolving constant. The subscripts p and d refer to perfect site
calcite offer the potential for better understanding and and dislocation site Gibbs free energies, respectively.
quantifying these surface site influences. Schott et al. (1989) pointed out that dislocations can
Because the concentration of highly reactive sur- influence dissolution kinetics (Fig. 7) by increasing the
face sites can be related, at least in part, to crystal potential energy of the solid (acrystal in Eq. (29)) and
defect density, Schott et al. (1989) used strained increasing the number of dislocation surface sites
calcites to investigate the rate dependence of calcite (term 2 in Eq. (30)). They found that even at high
dissolution on defect density. Following the work of dislocation densities (1011 cm 3) only about a 4%
Fig. 7. Schematic illustration of parallel process involved in crystal dissolution. Horizontal length of arrows is relative to rate and thickness
represents quantity of material dissolved. (After Schott et al., 1989.)
J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84 63
increase in activity occurred, so that the second (sur- model. They found that the following reactions pro-
face dislocation) influence was strongly dependent on vided an adequate description of calcite dissolution
reaction kinetics. Eqs. (29) and (30) were then com- and precipitation in dilute solutions (> denotes a
bined and modified to produce Eq. (31), where Stot is surface complex):
the total surface area and Sd is the area covered by
k1
defect sites (Stot can be used since SpHSd), and kp and > CO þ þ
3 þ 2H U > Ca þ H2 CO3 ð32Þ
k5
kd are the dissolution rates at perfect and dislocation
sites, respectively.
k2
Y n > Caþ þ H2 CO3 U > CO3 H0 þ CaHCOþ
3 ð33Þ
k6
R¼ ai i ðStot kp þ Sd kd Þ ð31Þ
i
k3
For unstrained calcite, Schott et al. (1989) found good > CO3 H0 þ CaHCOþ
3 U
k7
agreement with Busenberg and Plummer (1986) and
þ
Sjöberg and Rickard (1984a). However, for strained > Ca þ H2 CO3 þ CaCO03 ð34Þ
calcite with defect densities greater than about 107 to
108 cm 3, a major (2 – 3 times) increase in near k4
CaCO3ðsolidÞ U CaCO03 ð35Þ
equilibrium dissolution rates was observed. The fact k8
that little influence of defect density on dissolution
rates was observed at high undersaturations, where The overall reaction rate is given by
transport control dominates, confirmed that the in- 2
Rate ¼ k1 > CO
3 ðaHþ Þ þ ðk2 k5 Þ
crease in rates near equilibrium, where reaction rate
is surface controlled, was the result of an increase in > Caþ aH2 CO* þ k4 ðk6 k3 Þ
surface dislocations. 3
0
> CO3 H aCaHCOþ3 k7
3.1.5. The surface complexation model
One of the most interesting advances in under- > Caþ aH2 CO* aCaCO03 k8 aCaCO03 ð36Þ
3
standing calcite dissolution kinetics in simple solu-
tions came from the surface complexation model for 3.1.6. Direct observations of dissolving surfaces
carbonate minerals developed and initially applied to In addition to the TEM work by Schott et al. (1989)
calcite dissolution kinetics by Van Cappellen et al. described above, observations of dissolving calcite
(1993). The basic idea of their work is that the surfaces have been made directly (either in situ or ex
speciation of surface sites is strongly influenced by situ) using a variety of atomic or near-atomic scale
solution chemistry other than simple saturation state. techniques, including atomic force microscopy and
Just as the ratios of different carbonic acid species optical interferometry. This recent work has focused
vary as a function of PCO2 at a given pH, so can the largely on the growth of etch pits, their morphology
ratios of various surface complexes. At high PCO2 and distribution. There are significant differences in
values, the dissolution rate at a given saturation state, experimental conditions (e.g., pH, PCO2, ionic strength,
can be increased by the formation of carbonate com- Ca2 + concentration) among these various efforts, giv-
plexes with surface calcium ions. It was thus possible ing rise to obvious variation in computed dissolution
to correlate the dissolution kinetics of calcite with the rates; moreover, the resolution of the dissolution rate
relative abundances of surface complexes formed at itself in familiar units can be a complex matter.
the mineral – water interface. However, a common observation has been the distinct
Subsequently, Arakaki and Mucci (1995) were able differences in step velocities of opposed steps that
to arrive at a general equation for calcite dissolution form the etch pit. Two of the neighboring, intersecting
and precipitation in simple solutions by combining steps forming a corner are ‘‘slow’’ while the diagonally
their own experimental data with those of others opposite pair is ‘‘fast’’ (see Fig. 8). The difference in
(Plummer et al., 1978; Chou et al., 1989) and the velocities ranges from 2 to 4 (Park et al., 1996). The
Van Cappellen et al. (1993) surface complexation relative step velocities obviously determines the over-
64 J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84
Fig. 9. Sequence of 3D rendered images of scanning interferometer data from the 101̄4 calcite cleavage surface dissolving at pH f 9, PCO2
10 3.5. Panel (A) is the pristine surface; panels (B), (C) and (D) record development and coalescence of etch pits after 15-, 60-, and 240-min
reaction (Ertan et al., unpublished data).
areas of high (or maximum) step density; surrounding 5.4 nm s 1), although the description of pit depths is
terraces (‘‘intersectional’’ regions) are characterized by a function of instrumentation: shallow vs. ‘‘deep’’ pits
low step density. By calculating a step density from described by Liang et al. (1996a) are 0.3 vs. f 12 nm
etch pit wall angles, the authors compute the rate of for MacInnis and Brantley (1992); certainly depths of
deep etch pit dissolution to be 1.5 F 0.3 10 10 mol less than one nm would not have been observable by
cm 2 s 1 (cf., e.g. Busenberg and Plummer 1986, the optical system of MacInnis and Brantley (1992).
2.4 10 10 mol cm 2 s 1 at pH 5.60, PCO2 = 0). Conversely, the 800-nm etch pit depths measured after
The summed velocity of Liang et al. (1996a; 4.9 30 min of reaction by the latter workers would have
nm s 1) is comparable to the lateral velocity of etch been difficult to resolve with (high-resolution) AFM
pit walls observed by MacInnis and Brantley (1992; instruments.
66 J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84
This highlights a central problem in using AFM in (Dove and Platt, 1996). One alternative way of
dissolution studies; namely scale. Although AFM approaching this problem is through application of
provides high specific detail (permitting the ang- vertical scanning interferometry, which allows quanti-
strom-scale resolution of lateral step migrations), fication of both the evolution of surface features in
relating the velocity of step directions to the bulk three dimensions, and a relatively simple means of
dissolution rates described above requires several computing bulk rates. As an illustration, a sequence of
pieces of information: interferometer-derived images is shown in Fig. 9.
These depict 3D rendered data collected over the
1. a geometric model that relates (observable) course of calcite dissolution at a pH of f 9 and PCO2
motion of step traces to the three dimensional of 10 3.5. These data show the initial development of
removal of surface material at etch pits; etch pits on a 101̄4 cleavage surface, whose growth
2. an estimate of the statistical distribution and development ultimately lead to widespread inter-
(density) of etch pits across the reactive ference and coalescence. Through the use of fixed
surface; (unreacted) reference points, absolute differences in
3. and, most importantly, an understanding of surface height data can be computed over each time
how this process evolves during the approach step, leading to a detailed history of reaction rates in
to a ‘‘steady state’’ rate of removal. conventional units (mol/unit area/unit time).
Interferometry can also establish the statistical
Also, removal of material by step migration may distribution of rates (Lüttge et al., 1999), and thus
also occur at flat terraces, whose specific rate will be permits quantification of the overall heterogeneity of
far less than that observed at etch pits but whose a real the dissolution process. Recent dolomite dissolution
contribution to the total rate may be significant. Since experiments using scanning interferometry (Lüttge
etch pits grow both by widening and deepening, they et al., 2002, described below) have also highlighted
must ultimately coalesce (as shown by MacInnis and relationships between geometric and BET surface
Brantley, 1992; a model for etch pit size distributions areas, intrinsic and average dissolution rates.
was introduced by MacInnis and Brantley, 1993), and In a series of unique experiments, Shiraki et al.
the overall rate is thus a reflection of competing (2000) compared calcite dissolution rates measured by
processes. These are etch pit deepening; step migra- reacted fluid chemistry as well as by AFM step veloc-
tion, whose velocity reflects crystallographic control ities. AFM rates (given in mol cm 2 s 1) are some-
(Liang et al., 1996a,b) and pit depth (thus, the change what depressed relative to those determined from
in pit wall angle with time observed by MacInnis and solution chemistry; overall, however, there is fair
Brantley, 1992); coalescence of etch pit boundaries. agreement (Fig. 10). Step velocities also differed from
This results in a surface of variable relief containing those recorded in other AFM studies by a factor of f 3.
etch pit floors, step edges, and slowly reacting terraces. For example, their average velocity at a pH of 8.9 is 1.6
The long-term evolution of this process is at present nm s 1, less than that recorded by MacInnis and
not understood (long-term AFM experiments are cur- Brantley (1992; 5.4 nm s 1) or Liang et al. (1996a;
rently technically difficult), and requires a quantitative, 4.9 nm s 1). Rates measured by Shiraki et al. (2000)
comprehensive model to relate atomic-scale processes are also in good agreement (within error, see Fig. 10)
to bulk phenomenological observations. with the measured rates of Plummer et al. (1978).
In addition, these problems relate to current ques-
tions regarding the relationship between total surface 3.2. Influence of temperature
area, as quantified, e.g., by gas absorption techniques
(BET, Brunauer et al., 1938) and reactive surface area Although the influence of pressure on carbonate
(White and Peterson, 1990; Rufe and Hochella, 1999). reaction kinetics has not been investigated, several
This relationship is an area of active research in the studies have focused on how calcite dissolution kine-
field of mineral – solution kinetics (e.g., Gautier et al., tics vary with temperature. Early work on the topic
2001). Direct observation of atomic-scale processes was summarized by Plummer et al. (1979). This data
must ultimately relate to ‘‘whole mineral’’ reactivity set was confined to dissolution in acidic solutions
J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84 67
Fig. 10. Comparison of AFM-derived calcite dissolution rates with those from previous studies using powders. Closed squares with error bars
are AFM rates, closed triangles are rates from reacted solution chemistry, both taken from Shiraki et al. (2000); open square is datum from
MacInnis and Brantley (1992); open circles and solid line are data from Sjöberg (1976); dotted line is from the model of Chou et al. (1989); and
dashed line is from model of Plummer et al. (1978).
where transport control is generally rate limiting (see results are summarized in Fig. 3. The boundaries
previous discussion). A fairly wide range (about 8 – 60 between H + -dependent, transitional and H + -inde-
kJ mol 1) of activation energies was reported. How- pendent regimes were found to move to lower pH
ever, probably the most reliable values in dilute values with increasing temperature.
solutions were close to the same; 8.4 kJ mol 1 by Gutjahr et al. (1996a) found a high activation
Plummer et al. (1978) and 10.5 kJ mol 1 by Sjöberg energy (35 kJ mol 1) in the near equilibrium region
(1978). This is a relatively small change in dissolution confirming that the calcite dissolution reaction is sur-
rate with temperature of about 13% for every 10 jC face controlled in this region of saturation. Interest-
(e.g., rates of biologic processes typically roughly ingly, they also found that the reaction order changed
double for each 10 jC change in temperature). from 2 at 20 jC to 1.2 at 70 jC, whereas the rate
Subsequently, Sjöberg and Rickard (1984b) made a constant increased from 1 10 10 to 2 10 10 mol
much more in depth study of the influences of temper- cm 2 s 1.
ature on carbonate dissolution kinetics under a variety Outside of this limited amount of work and that in
of conditions. Although these conditions were con- seawater (see next section), little is known about the
strained to inhibitor-free solutions far from equili- influences of temperature on calcite dissolution
brium, they still revealed a quite complex impact of kinetics in the surface-controlled near equilibrium
temperature on calcite dissolution kinetics. Their region that is common in natural systems or how the
68 J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84
influence of inhibitors may interact with changing Since, as expected, it has been observed that inhib-
temperature. This imposes a very substantial limita- itory influences often dramatically increase with
tion on our ability to predict calcite dissolution increasing saturation state and there is site competition
kinetics in the ‘‘real world’’. among inhibitors and between inhibitors and major
ions, results are often contradictory and not readily
3.3. Influences of solution composition and reaction comparable. Therefore, only relatively few of these
inhibitors studies, which illustrate major points, will be dis-
cussed.
3.3.1. General considerations Because the inhibitory influence of ions occurs
The previously discussed studies of calcite in dominantly by their interactions with the calcite sur-
compositionally simple and generally dilute solutions faces, it is necessary to understand how these ions
provide information that is primarily of interest from a interact with surfaces. This is a very major and com-
‘‘pure chemistry’’ standpoint, but of limited utility for plex topic, and calcite surface chemistry will, there-
most natural waters which are chemically complex fore, only be dealt with to a rather limited degree in this
and where reactions often proceed relatively close to paper. It is worth noting here, however, that basically
equilibrium (e.g., the deep ocean). However, just as two approaches are used in discussing surface inter-
equilibrium thermodynamics provide a valuable start- actions of ions that inhibit growth or dissolution
ing point for looking at deviations from equilibrium in kinetics (for references and discussion relative to
natural systems, so also do the results of studies of carbonates, see Gutjahr et al., 1996b). The first is the
calcite dissolution in ‘‘pure’’ systems provide a basis Langmuir – Volmer model that assumes reversible
for looking at the influences of inhibitors common in adsorption at specific surface sites according to the
natural waters. Langmuir adsorption isotherm. The second is the
The very major impact that ‘‘foreign’’ ions can Cabrera and Vermilyea model that assumes irreversi-
have on calcite dissolution was recognized by early ble adsorption of additives on a terrace whereby the
investigators such as Weyl (1958) and Terjesen et al. advancing steps get caught by additives while they
(1961), who demonstrated that metal ions even at sub- grow between them (seen also in AFM work; see, e.g.,
micromolar concentrations can severely retard calcite Dove and Hochella, 1993; Gratz et al., 1993). Exam-
dissolution. Table 1 gives a listing of studies of in- ples of these two approaches follow later in this
hibitors of calcite dissolution. Unfortunately, many of section. It is important to note that generally the
these studies have been done under very different kinetics of the adsorption reaction are ignored and that
saturation states, and in solutions often not containing only the initial linear part of a Langmuir adsorption
the same major ions and using differing techniques. isotherm is considered. It has usually not been dem-
Table 1
Studies of calcite dissolution inhibitors
Inhibitor References
Ca2 + Sjöberg (1978), Buhmann and Dreybrodt (1987)
Mg2 + Akin and Lagerwerff (1965), Benjamin et al. (1977), Berner (1967, 1975, 1978), Buhmann and Dreybrodt
(1987), Gutjahr et al. (1996b), Möller (1973), Möller and Rajagopalan (1975), Plath et al. (1980), Plummer and
Mackenzie (1974), Reddy and Wang (1980), Sjöberg (1978), Walter and Morse (1984a), Weyl (1958, 1965)
Sr2 + and Ba2 + Gutjahr et al. (1996b)
Transition and Gutjahr et al. (1996b), Nestaas and Terjesen (1969), Salem et al. (1994), Terjesen et al. (1961)
heavy metals
Sulfate Akin and Lagerwerff (1965), Buhmann and Dreybrodt (1987), Erga and Terjesen (1956), Kushnir (1983),
Mucci et al. (1989), Sjöberg (1978)
Phosphate Berner and Morse (1974), Morse (1974), Nancollas et al. (1981), Reddy (1977), Sjöberg (1978),
Svensson and Dreybrodt (1992)
Silica and nitrate Morse (1974)
Organics Barwise et al. (1990), Berger (1967), Compton and Sanders (1993), Morse (1974), Suess (1970)
J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84 69
onstrated that beyond some inhibitor solution con- and pure NBS calcite could largely be explained in
centration the dissolution rate becomes independent terms of the influence of phosphate associated with the
of inhibitor concentration as would be predicted natural samples.
for surface site saturation by the Langmuir– Volmer Following on the early work of Terjesen et al.
model. (1961) and Nestaas and Terjesen (1969) that demon-
strated a strong inhibitory effect of trace metals on
3.3.2. Inhibitor behavior in relatively dilute solutions calcite dissolution, Salem et al. (1994) studied the
In the previous discussion of calcite dissolution inhibitory influences of several transition metals on
kinetics, the primary variables influencing saturation calcite under diffusion controlled conditions in dilute
state and reaction kinetics were parameters such as solutions. They observed major reductions in reaction
PCO2 and pH in the carbonic acid system, the anion rates even at micromolar concentrations of the metals
variable. However, it has also been found that the and found the relative order of inhibitory strength to
cation, Ca2 +, concentration can also influence reac- be Ni > Cu > Mn > Co. They attributed the inhibitory
tion rate beyond its simple influence on saturation. effects of these metals to Langmuir adsorption that
Sjöberg (1978) found about a 17% increase in the caused the blockage of active surface dissolution sites.
rate constant over a range of initial calcium concen- More recently, Lea et al. (2001) have also documented
trations from 0 to 10 mM which could be interpreted significant reductions in AFM step velocities with
in terms of a Langmuir-type adsorption isotherm. A additions of Mn2 + greater than 0.5 AM.
similar, but substantially smaller influence, was found There have been surprisingly few studies of the
for magnesium whose influence increased with influence of organic compounds on the dissolution of
increasing magnesium concentration and saturation calcite in dilute solutions. Barwise et al. (1990) found
state. The combined influences of Ca2 + and Mg2 + by studying etch pit development that the di-anions of
were investigated and a complex relationship was maleic and tartaric acids are powerful inhibitors capa-
found leading Sjöberg to conclude that it would be ble of completely stopping proton-induced calcite
impractical to produce a general equation describing dissolution. Compton and Sanders (1993) made the
the effect of inhibitors on calcite dissolution rates. interesting observation that in equilibrated acidic
Sjöberg (1978) also found that both phosphate and solutions, humic acids had no influence on calcite
sulfate could inhibit calcite dissolution and that their dissolution rates, but that if fresh sodium salts of
influences increased with increasing calcium and humic acid were added significant inhibition of dis-
magnesium concentrations. This is probably the result solution occurred.
of, at least in part, the adsorption of calcium and
magnesium that produces a more positive and cationic 3.3.3. Studies of reaction kinetics and inhibitors in
surface (cation bridging) that increases the adsorption seawater
of phosphate and sulfate. Sjöberg (1978) also was able Probably the most historically important studies of
to demonstrate that for phosphate-free artificial sea- calcite dissolution were not conducted in the labora-
water, rates could be well predicted by simply adjust- tory, but rather in the ocean by Berger (1967) and
ing the rate constant for the presence of sulfate. His Peterson (1966), who hung packets containing fora-
results showed that the other major and minor ions of minifera tests and polished calcite spheres, respec-
seawater have little influence on calcite dissolution tively, at different depths in the Pacific Ocean. Their
kinetics. rates of dissolution, based on weight change, were
More recently, Buhmann and Dreybrodt (1987) then related to the in situ saturation states of the
found that for diffusion-controlled calcite dissolution, seawater. The most important finding was that the
the inhibitory influences of calcium, magnesium and dissolution rate of calcite in seawater is not directly
sulfate could largely be accounted for by the ‘‘common proportional to the extent of undersaturation and
ion effect’’, that results in changes in the differences therefore simple first order kinetics cannot be used
between surface and bulk solution concentrations of to describe this process. Similar results were later
Ca2 +. Svensson and Dreybrodt (1992) found that the obtained for the Atlantic Ocean (Milliman, 1977;
difference in behavior between natural calcite samples Honjo and Erez, 1978). Their findings are in general
70 J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84
Fig. 13. The ‘‘flow diagram’’ for studies of inhibitors of calcite dissolution in seawater. (From Morse, 1974.)
constant, but a very large influence was observed on and controversial field of inquiry. However, at some
the reaction order which increased up to about 16 at a point these studies will need to bridge the gap between
phosphate concentration of only 10 AM (Fig. 14). the carefully controlled laboratory investigations dis-
These results were interpreted by Berner and Morse cussed here and the complex natural oceanic environ-
(1974) in terms of the previously discussed Cabrera ments. At present it is clear that there is a long way to
and Vermilyea model. Berner and Morse (1974) go in doing so. This is exemplified by the recent paper
hypothesized that increasing concentrations of phos- by Hales and Emerson (1997) where simple first order
phate at active sites (kinks) resulted in a closer kinetics, with a vastly lower rate constant than ob-
spacing of phosphate ions that progressively raised served in experiments, are necessary to model their
the ‘‘critical undersaturation’’ for dissolution on retreat- results for fluxes from deep sea sediments.
ing steps.
A large number of papers have been written con-
cerning the dissolution of biogenic carbonates and 4. Dissolution of other sedimentary carbonate
calcium carbonate-rich sediments, speculating on deep minerals
sea carbonate dissolution kinetics based on observa-
tions of fluxes from sediments and microelectrode 4.1. Aragonite dissolution kinetics
measurements in sediments. The topic of carbonate
accumulation, in both shallow and deep water sedi- Aragonite is by far the dominant carbonate mineral
ments, and the role of biologic processes, both in the in tropical and subtropical shallow water carbonate-
water column and in sediments, remains a vigorous rich sediments. It is found in deep sea sediments
72 J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84
contain absolute reaction rates. They are chiefly con- on dissolution rate alone. Thorstenson and Plummer
cerned with relationships between solid phase stoichi- (1977) approached this problem by introducing the
ometry, bulk solution composition, and the relative concept of stoichiometric saturation. For a magnesian
rate of magnesium vs. calcium release to solution. calcite having a composition Ca(1 x)MgxCO3, they
Land (1967) observed that short-term (24 h) dissolu- proposed that the equilibrium constant is given by the
tion of skeletal magnesian calcites yielded congruent product of activities
release to solution; by comparison, longer term experi-
ments produced incongruent enrichment of dissolved IAPmag cal ¼ ðCa2þ Þ1x ðMg2þ Þx ðCO2
3 Þ, ð37Þ
magnesium at 25 jC, with the relative rate of this
where x = mole fraction lattice magnesium. Magnesian
process proportional to the Mg content of the reactant
calcite is thus assumed to react as a one-component
calcite. Dolomite was not observed to precipitate at
phase having a fixed composition. With this assump-
25 jC, but did exsolve under hydrothermal conditions
tion, the Gibbs – Duhem equality of chemical potentials
(300 jC), and the relative rate of this process (all other
of CaCO3 and MgCO3 between solid and aqueous
things being equal) was also roughly proportional to
phases can no longer be asserted. However, stoichio-
solid reactant Mg content (Land, 1967).
metric saturation was offered as a resolution to the
The problems regarding magnesian calcite that
problem of magnesian calcites that react so slowly as to
bear on their dissolution kinetics can be organized
preclude thermodynamic equilibrium, giving the
as three general questions:
appearance of reacting with apparently fixed compo-
sition. This concept was controversial at the time of its
1. What is the appropriate formulation of the
introduction (see responses by Garrels and Wollast,
relationship between the activity of solution
1978; Lafon, 1978; Thorstenson and Plummer, 1978)
components (calcium, magnesium, carbonate
and its usage has not been fully resolved. Plummer and
ion, etc.) and saturation state with respect to a
Mackenzie (1974), in earlier free drift experiments
calcite of a given magnesium content?
dissolving biogenic magnesian calcite at fixed PCO2,
2. How can apparent differences in reactivity be
derived equilibrium constants pffifor the dissolution rate
related to the phase’s origin (synthetic or
from reciprocal root time (1/ t) plots of free drift data
inorganic vs. biogenic)? and
(see below). Their approach implicitly assumed that
3. How can reactive surface area be quantified in
only the CaCO3 component attains a true saturation
biogenic magnesian calcites?
state, represented by
There have been several review volumes (e.g., Mack-
enzie et al., 1983; Morse and Mackenzie, 1990) that IAPCaCO3 ¼ ðCa2þ ÞðCO2
3 Þ: ð38Þ
have covered this ground in extensive detail; what
follows here is a brief encapsulation of aspects rele- However, subsequent solubility determinations have
vant to dissolution kinetics. demonstrated the utility of the concept of stoichiomet-
ric saturation, and Walter and Morse (1984a) have
4.2.1. Stoichiometric saturation shown that magnesian calcite reaction orders com-
In the case of magnesian calcites, the practice of puted from rate vs. (1 X)stoich-sat relations are not
relating reaction rates to the distance from equilibrium vastly dissimilar to those of pure calcite.
(X) becomes problematic, and controversy has In still earlier work, Land (1967) observed that
revolved around just how this equilibrium should be dissolution of biogenic magnesian calcite in pure
defined. This problem arises partly from uncertainties water proceeded incongruently, which he attributed
in free energy values, from the larger question of to a faster dissolution rate of MgCO3 vs. CaCO3
whether magnesian calcites ever attain a true meta- component. Plummer and Mackenzie (1974) argued
stable equilibrium state, and how to represent this that the ratio of Mg2 + /Ca2 + in solution reflected ini-
relationship in thermodynamic calculations. The issue tially congruent dissolution, but that the dissolving
of how magnesium substitution affects solubility has phase differed in composition (and thus, presumably,
thus received far more attention than that of the effect solubility) from the bulk solid (initial = 24.7 vs.
74 J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84
bulk = f 18 mol% MgCO3). Thus, the phase govern- that microstructural differences in biogenic magne-
ing the initial dissolution rate in a multimodal assem- sian calcites exert a far stronger influence on their
blage of magnesian calcite mineralogies is the most dissolution kinetics than their relative thermody-
soluble. Plummer and Mackenzie (1974) argued that namic stabilities.
incongruent dissolution follows a so-called parabolic Walter and Hanor (1979) found orthophosphate to
rate law, which casts time (but not concentration) as inhibit the dissolution rate of biogenic magnesian
the independent variable, and derived rate constants calcite in a manner similar to that observed for pure
based on the following relationship, calcite, i.e., the inhibition increased with increasing
phosphate concentration and as calcite saturation was
Ci ¼ 2ð1 ni Þk t 1=2 ð39Þ approached. Biogenic magnesian calcites also show
larger unit cell volumes, and do not show smooth
In this equation, Ci is the dissolved concentration of variation in c/a axial ratios with magnesium substitu-
the ith component in bulk solution, k is the rate tion compared to synthetic phases (Bischoff et al.,
constant, t is time, and the term (1 ni) is a correction 1983). This difference has been attributed to greater
factor to accommodate the stoichiometry of the incon- carbonate ion positional disorder (Bischoff et al.,
gruent reaction (ni = the ratio of stoichiometric coeffi- 1985; Paquette and Reeder, 1990). Bischoff et al.
cients of component i in mineral product over mineral (1987) concluded that biogenic phases, when com-
reactant). pared to synthetic magnesian calcites of similar Mg
In similar free drift experiments, Wollast and Rein- content, are more soluble, and moreover do not show
hard-Derie (1977) confirmed these results, as well as the smooth variation in stability vs. composition
subsequent reprecipitation of a calcite poorer in Mg characteristic of the latter. They proposed that the
than the initial reactant. Bertram et al. (1991) explored use of a single IAP vs. solid composition curve was
the effect of temperature on dissolution of synthetic inappropriate, and that subsequent work should con-
magnesian calcites (f 2 to 19 mol% MgCO3), finding sider the use of three curves that reflect the inherent
solubility to decrease with temperature in a manner heterogeneity of the total dataset: synthetic materials,
similar to pure calcite. carefully treated biogenic materials, and untreated
Using pH-stat techniques, Walter and Morse biogenic materials.
(1985) measured dissolution rates of carefully treated
biogenic magnesian calcites (11– 18 mol% MgCO3), 4.2.3. The role of reactive surface area
computing rate constants and reaction orders fit to Eq. In comparing the dissolution rates of pure calcite
(3), calculating saturation state from stoichiometric and biogenic magnesian calcites having a range of Mg
activity products and stoichiometric constants (Walter content, Walter and Morse (1984b, 1985) observed
and Morse, 1984a). Rates were reported on a per unit that the primary differences were most strongly con-
mass (Amol g 1 h 1) vs. per unit area basis because trolled not by magnesium composition but by relation-
of large differences in reactive vs. BET surface area ships between total (BET) surface area, grain size and
(see below). They observed no systematic variation in microstructure. They observed that although dissolu-
rate parameters with MgCO3 content, although reac- tion rates in general increased with BET surface area,
tion orders are somewhat higher (n = 3.2– 3.5) com- these increases did not show a simple proportionality.
pared to low-Mg calcite. By assuming a unit ratio of reactive to total surface
area in synthetic calcite (f 5-Am grain diameter), they
4.2.2. Biogenic vs. abiogenic phases were able to show that the complex microstructures
In comparing calcites of similar Mg content, characteristic of biogenic Mg calcites result in their
biogenic calcites exhibit greater structural and chem- having reactive surface areas that constitute as little as
ical heterogeneity and accommodate higher concen- 1% or less of the total area. This must in part result
trations of nonstoichiometric components such as from the large differences in transport efficiency of
water and hydroxyl (Mackenzie et al., 1983; Gaffey, gases vs. dissolved solutes, and from the high tortuos-
1988), sodium, and bicarbonate. As is discussed ity of complex networks of small pore spaces charac-
below, a key observation made in these papers is teristic of skeletal pore systems. Thus, in these cases,
J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84 75
BET area measurements are virtually useless as a sedimentary dolomites of Recent age that have not
measure of reactive area. undergone extensive recrystallization. Although it is
In summary, magnesian calcites exhibit unique difficult to separate the independent contributions of
properties that complicate their reaction kinetics, and disordering and excess calcium, Chai et al. (1995)
the actual dissolution mechanisms are to a large extent have shown that even small amounts of calcium
still unresolved. In synthetic phases, the addition of substitution generate positive heats of formation for
lattice magnesium below about 6 mol% MgCO3 has a dolomite, and thus are strongly destabilizing. Inor-
relatively minor effect on thermodynamic stability, and ganic calcite, in contrast, can accommodate a compa-
may actually decrease solubility at f 3 mol%; from 6 ratively large extent of magnesium substitution, with
mol% up to f 15 mol%, there is a well-documented only minor changes in enthalpy (Navrotsky and Capo-
overall increase in solubility (Mucci and Morse, 1984; bianco, 1987). This limited understanding of dolomite
Bischoff et al., 1987; Navrotsky and Capobianco, solubility thus handicaps the description of near-equi-
1987; Bischoff, 1998). Because of the accessibility librium kinetics, whose expressions typically include
and importance of biogenic magnesian calcite in an explicit free energy term. To be complete, a model
natural marine environments, dissolution work has of the kinetics of dolomite dissolution (and growth)
focused predominantly on these materials. However, must ultimately address the large difference in reac-
correct comparison of dissolution rates of biogenic tivity compared to calcite as well as magnesite, and the
magnesian calcites depends strongly on evaluation of role played by crystal structure and composition.
reactive surface area, and differences in dissolution Busenberg and Plummer (1982) undertook the first
rate as a function of Mg content alone may be minor by significant laboratory measurement of dolomite disso-
comparison. As a practical device, the use of stoichio- lution rates at pH > 2 and at temperatures typical of
metric saturation to estimate distances from equili- sedimentary environments. Actual dissolution rates
brium allows satisfactory comparison of phases of were measured by the weight loss of cleavage frag-
varying composition, but is not free of uncertainty ments, and conditions included large variations in fixed
on a theoretical basis. Lastly, it is unclear what ‘‘net’’ PCO2 (0 to f 1 bar), pH ( < 1 to f 10), solution com-
effect these various physical properties (initial Mg position (added calcium and magnesium), and temper-
content, distribution, and heterogeneity, grain size ature (5 –65 jC). Because this paper has become the de
and microstructure, defect density, solution composi- facto point of comparison in all subsequent work, and
tion) have on the reaction path (congruent or incon- because of the limited literature on dolomite dissolu-
gruent) of magnesian calcites during stabilization in tion, the work is described below in some detail.
natural environments (Bischoff et al., 1993). Following the earlier approach to calcite of Plummer
et al. (1978); Busenberg and Plummer (1982) proposed
4.3. Dolomite dissolution kinetics that their rate data could be fit to the expression,
1 1 1
(and PCO2 f 0), a third term describes the dissolu- Because these reactions consume only one half of the
tion rate’s approach to a constant value. Because molar equivalent of dolomite per mole reactant (H +,
maintaining moderate pH values (]4.5) at a high H2CO03, H2O), they concluded this was the origin of
PCO2 requires the addition of carbonate alkalinity, the the observed reaction order n = 1/2. (This conclusion
reduction in observed vs. predicted rates was inter- does not, however, explain the initial observed in-
preted to be the influence of HCO3 ion, described congruent step, in which magnesium is released prior
by the addition of a fourth ‘‘backward’’ (precipita- to calcium).
tion) rate term in Eq. (40). Rate constants at 25 jC Busenberg and Plummer (1982) observed the order
for k1 through k4 are 1.7 10 8, 4.3 10 10, with respect to hydrogen ion to increase with temper-
2.5 10 13, and 2.8 10 8 (mol cm 2 s 1), ature above 45 jC, giving values of 0.6 at 55 jC and
respectively. 0.7 at 65 jC. In earlier rotating disk experiments
Busenberg and Plummer (1982) also observed a (pH < 2), Lund et al. (1973) described a similar varia-
complex pattern of incongruent dissolution behavior. tion of reaction order with respect to HCl concentra-
In dissolution runs using pure water, the initial ratio of tion and temperature (n = 0.44 at 25 jC, 0.61 at 50 jC,
Ca to Mg released to solution was less than unity and 0.85 at 100 jC). In experiments using a fluidized
(leaving the solid calcium-rich). This pattern then bed reactor, steady state rates measured by Chou et al.
reversed itself, with the solid yielding greater calcium (1989) are similar to those of Busenberg and Plummer
than magnesium (leaving a magnesium-rich surface), (1982) in the region of high pH (8 – 9). However,
but with sufficient reaction the yield approached increasingly higher rates were measured at lower pH,
congruence. Busenberg and Plummer (1982) used this resulting in an overall disagreement of slightly less
evidence of incongruent surface reactions to construct than an order of magnitude at pH 3. Chou et al. (1989)
a rather elaborate reaction scheme consistent with Eq. did not fix the PCO2 prevailing within the reactor as did
(40). The parallel reactions associated with k1, k2, and Busenberg and Plummer (1982), but computed PCO2
k3 are from the alkalinity and pH of the reacted solution. In
k10 addition to the disagreement at low pH, Chou et al.
CaMgðCO3 Þ2 þ Hþ ! MgCO3 þ Ca2þ þ HCO
3 (1989) also calculated a reaction order using Eq. (40)
ð41Þ of n = 3/4. Rate constants at 25 jC for k1 through k3 are
CaMgðCO3 Þ2 þ H2 CO03 2.6 10 7, 1.0 10 8, and 2.2 10 12. Chou et al.
(1989) argued that the scheme of successive reactions
k20 described by Busenberg and Plummer (1982) would
! MgCO3 þ Ca2þ þ 2HCO
3 ð42Þ
not produce fractional orders less than unity, and
instead attributed their observed order to the degree
CaMgðCO3 Þ2 þ H2 O
of surface protonation and the availability of surface
k30
complexes.
! MgCO3 þ Ca2þ þ HCO
3 þ OH
ð43Þ Relatively recent dissolution work at pH < 5 was
presented by Gautelier et al. (1999). In these rotating
in which ‘‘MgCO3’’ is the Mg-enriched surface. In disk experiments, log rates were found to be linear
order to accommodate the evidence for incongruent with respect to pH over the range of 1 – 5, giving
reactions, they argue that the subsequent reactions k1 = 9.8 10 8, 5.1 10 7, and 1.8 10 6 mol
(listed also respectively) involving ‘‘MgCO3’’ are cm 2 s 1 and orders n = 0.63, 0.73, and 0.80 at
slower and rate-limiting: 25, 50, and 80 jC, both respectively. Below a pH of
k 1W
f 1, the dissolution rate approached a plateau and
MgCO3 þ Hþ ! Mg2þ þ HCO
3 ð44Þ became independent of pH. Gautelier et al. (1999)
used their measured pH and dissolution rate data to
k 2W estimate the diffusional boundary layer thickness and
MgCO3 þ H2 CO03 ! Mg2þ þ 2HCO
3 ð45Þ
surface hydrogen ion activity. They argued that a
k 3W decrease in the ratio of surface to bulk hydrogen ion
MgCO3 þ H2 O ! Mg2þ þ HCO
3 þ OH
ð46Þ activities (aH + surf /aH + bulk) with increasing bulk pH
J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84 77
indicates the increasing role of solute transport on (1999) are not independent of assumptions regarding
measured reaction rates (Gautelier et al., 1999, their DBL, and it is thus difficult to reconcile the extent to
Fig. 9, p. 23). They concluded that the fractional order which they reflect uncertainty in these assumptions as
of the reaction with respect to hydrogen ion is well as measured values. Moreover, the use of surface
responsible for this behavior. Fig. 15 compares the to bulk activity ratios in this case can be somewhat
data from Busenberg and Plummer (1982), Chou et al. deceptive, as the observed relationship may result from
(1989), and Gautelier et al. (1999) at 25 jC. a small but constant offset (error?) between estimated
Talman and Gunter (1992) found the order of surface and bulk activities. It is also difficult to
hydrogen ion dependence to increase with temper- evaluate to what degree either concentration or trans-
ature, but with increasing evidence of stirring rate port of CO2 away from the reacting surface may have
dependence at temperatures over 100 jC. By compar- influenced measured rates, given Busenberg and
ison, Lund et al. (1973) found no stirring rate (and thus Plummer’s (1982) observation of the rate’s depend-
no solute transport) dependence of dolomite dissolu- ence on CO2. Gautelier et al. (1999) also noted a
tion at temperatures below 50 jC (in 1M HCl). decrease in apparent activation energy with increasing
Busenberg and Plummer (1982) also found no disso- pH (46 kJ mol 1 at pH = 0 vs. 15 kJ mol 1 at pH 5).
lution rate dependence on stirring rate at 45 jC, pH An activation energy of 15 kJ mol 1 is characteristic
3 – 4.4. However, the conclusions of Gautelier et al. of diffusion-controlled reactions, but although this
Fig. 15. Dolomite dissolution rate data as a function of pH, taken from the following sources: open circles, Busenberg and Plummer (1982)
(weight loss); open squares, Gautelier et al. (1999) (rotating disk, solution chemistry); solid curve, equation of Chou et al. (1989) (column
reactor, solution chemistry); filled circle, Lüttge et al. (2000) (fluid cell, vertical scanning interferometer).
78 J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84
pattern is compatible with the authors’ assertion of Following Van Cappellen et al. (1993), they use the
increasing transport control with increasing pH, it may above equation and data from Busenberg and Plummer
also reflect the role of CO2 and its corresponding (1982) and Chou et al. (1989) to compute a second
temperature-dependent solubility. order dependence on > CO3H0 (n = 2), and conclude
In a recent series of papers, Pokrovsky and cow- that this is evidence for protonation of the two surface
orkers (Pokrovsky and Schott, 1999; Pokrovsky et al., carbonates adjacent to hydrated Ca and Mg sites prior
1999a,b) have championed the surface complexation to cation release. In like manner, they argue that under
model (see above discussion of Van Cappellen et al., neutral to high pH conditions, the decrease in the
1993) as a means of resolving the reactivity and overall rate (attributed to the action of the third term
reaction rates of dolomite and magnesite. Surface in Eq. (40)) can be understood as the replacement of
complexation constants in these papers are recovered > CaOH2+ and >MgOH 2+ by less protonated > CaOH0
by a variety of methods (surface potentiometric titra- and > MgOH0 sites (see also Van Cappellen et al.,
tion, streaming potential and electrophoresis) that in 1993):
turn involve a large number of assumptions and
empirical parameters that may defy simple compari- þ n
RH2 O ¼ kMe ðf>CaOHþ
2 and >MgOH2 gÞ ð48Þ
sons to existing data. At first appearance, the dissolu-
tion rate data for dolomite of Pokrovsky et al. (1999a),
most of which are recovered under relatively high pH with order n = 2. They also suggest that the systematic
(5.4 – 11), cannot be easily related to the overall trends variation of this reaction order in the carbonate min-
in log rate vs. pH in either Busenberg and Plummer erals (n = 1, 2, and 4, for calcite, dolomite, and mag-
(1982) or Chou et al. (1989), and differ in apparent nesite, respectively) indicates the increase in the num-
order. For example, rates measured by Pokrovsky et al. ber of metals adjacent to a carbonate group that require
(1999a) in solutions of pH of less than f 8 (and total hydration. In addition, Pokrovsky et al. (1999a) sug-
carbon < 0.001 M) show little change (judging by the gest that dissolved HCO 3 or CO3
2
will also titrate
apparent experimental uncertainty) with respect to pH. protonated metal sites to produce MeHCO30 and
Above pH 8, the large concentrations in total carbon MeCO 3 , and thus act to inhibit dissolution (Busenberg
(0.001 – 0.08 M) also complicate comparisons of their and Plummer’s ‘‘backward’’ reaction).
rates with earlier studies. Despite the insight provided by the above models
In spite of the difficulties in comparing earlier data, in terms of possible surface reaction mechanisms, a
Pokrovsky et al. (1999a) argue that their observed rates fundamental issue remains in the relationship between
can be correlated with surface speciation predicted the observed reaction rate and the extent to which the
from the complexation model. Three primary hydra- total surface participates in this process. In dissolution
tion sites are invoked for dolomite: > CO3H0, > M- experiments at pH 3, Lüttge et al. (2002) used vertical
gOH0, and > CaOH0. Depending on pH, these sites scanning interferometry (see above) to measure the
may also play host to the additional surface spe- development of etch pits on three different cleavage
cies > CO 3 Mg + , > CO 3 Ca + , > CO 3 , > MgOH 2 ,
+
surfaces. They were able to compare the rates of etch
0 +
> MgHCO3, > MgCO3 , > MgO , > CaOH2, CaHCO3, pit growth with the dissolution rate of the surrounding
> CaCO
3 , and > CaO . Pokrovsky et al. (1999a) (flat) surface. Etch pit rates vary with the cleavage
ascribe the four terms in Busenberg and Plummer’s surface on which they develop, with two surfaces both
(1982) rate equation to reactions involving surface giving rates of 1 10 10 mol cm 2 s 1, and the
species. For example, Pokrovsky et al. (1999a) argue third producing etch pits that grew at rates up to 10 9
that the increase in dissolution rate at pH < 6 reflects mol cm 2 s 1. However, the rates with respect to
the increasing protonation of > CO 3 surface sites, etch pits are significantly faster than those measured
such that the first term in Eq. (40) becomes using the ‘‘whole’’ surface, which gave a mean of
1.08 10 11 mol cm 2 s 1 (see Fig. 15). These
rates are also significantly lower than the rate meas-
ured by Busenberg and Plummer (1982) at pH 3.0
RHþ ¼ kCO3 f> CO3 H0 gn ð47Þ ( PCO2 f 0) of 2.95 10 10 mol cm 2 s 1 (cf. also
J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84 79
kinetic scheme difficult. Lastly, despite its utility, the Barwise, A.J., Compton, R.G., Unwin, P.R., 1990. The effect of
representation of magnesian calcite as a phase of fixed carboxylic acids on the dissolution of calcite in aqueous solu-
tion. J. Chem. Soc., Faraday Trans. 86, 137 – 144.
stoichiometry results in a rate law in which thermody- Bathurst, R.G.C., 1975. Carbonate Sediments and Their Diagenesis,
namic and kinetic terms are poorly separated. As 2nd edn. Developments in Sedimentology, vol. 12. Elsevier,
current work on surface mechanism has been domi- New York, NY.
nated by pure calcite, magnesian calcites have also yet Benjamin, L., Loewenthal, B.L., Marais, G.V.R., 1977. Calcium
carbonate precipitation kinetics: II. Effect of magnesium. Water
to see any scrutiny from these newer approaches.
S. Afr. 3, 155 – 165.
In contrast to calcite, work on dolomite and mag- Berger, W.H., 1967. Foraminiferal ooze: solution at depths. Science
nesite dissolution kinetics is still in its infancy. Most 156, 383 – 385.
of the experimental work published thus far has been Berner, R.A., 1967. Comparative dissolution characteristics of car-
cast in the formalism of Plummer and coworkers, and bonate minerals in the presence and absence of aqueous magne-
requires much more attention to the near-equilibrium sium ion. Am. J. Sci. 265, 45 – 70.
Berner, R.A., 1975. The role of magnesium in the crystal growth of
region. The mechanistic meaning of the observed calcite and aragonite from seawater. Geochim. Cosmochim. Acta
variation in fractional order with respect to hydrogen 39, 489 – 504.
ion activity remains unresolved, as do observations Berner, R.A., 1978. Rate control of mineral dissolution under Earth
concerning the apparent reaction of MgCO3 and surface conditions. Am. J. Sci. 278, 1235 – 1252.
CaCO3 as independent components. Approaches that Berner, R.A., Morse, J.W., 1974. Dissolution kinetics of calcium
carbonate in sea water: IV. Theory of calcite dissolution. Am. J.
incorporate the surface complexation model of Van Sci. 274, 108 – 134.
Cappellen et al. (1993) appear promising, but the Bertram, M.A., Mackenzie, F.T., Bishop, F.C., Bischoff, W.D.,
extensive parameterization will require substantially 1991. Influence of temperature on the stability of magnesian
more verification. calcite. Am. Mineral. 76, 1889 – 1896.
Binnig, G., Quate, C.F., Gerber, C., 1986. Atomic force microscope.
Phys. Rev. Lett. 56, 930 – 933.
Bischoff, W.D., 1998. Dissolution enthalpies of magnesian calcites.
Acknowledgements Aquat. Geochem. 4, 321 – 336.
Bischoff, W.D., Bishop, F.C., Mackenzie, F.T., 1983. Biogenically
This research was supported in part by a grant from produced magnesian calcite: inhomogeneities in chemical and
the Department of Energy (DE-FG03-00ER15033; physical properties; comparison with synthetic phases. Am.
Mineral. 68, 1183 – 1188.
Morse) and the Louis and Elizabeth Scherck Chair
Bischoff, W.D., Sharma, S.K., Mackenzie, F.T., 1985. Carbonate
endowment (Morse). We would like to dedicate this ion disorder in synthetic and biogenic magnesian calcites: a
paper to the memory of Dr. Lennart Sjöberg who died Raman spectral study. Am. Mineral. 71, 581 – 589.
in 1996. His outstanding work, in collaboration with Bischoff, W.D., Mackenzie, F.T., Bishop, F.C., 1987. Stabilities of
Dr. David Rickard, was a major and lasting contribu- synthetic magnesian calcites in aqueous solution: comparison
with biogenic materials. Geochim. Cosmochim. Acta 51, 1413 –
tion to our understanding of carbonate dissolution
1423.
kinetics. We are grateful for the many helpful Bischoff, W.D., Bertram, M.A., Mackenzie, F.T., Bishop, F.C.,
comments and suggestions on the manuscript by Drs. 1993. Diagenetic stabilization pathways of magnesian calcites.
Gerald M. Friedman, Fred T. Mackenzie and Alfonso Carbonates and Evaporites 8, 82 – 89.
Mucci. Blum, A., Lasaga, A., 1987. Monte Carlo simulations of surface rate
laws. In: Stumm, W. (Ed.), Aquatic Surface Chemistry: Chem-
ical Processes at the Particle – Water Interface. Wiley, New York,
NY, pp. 255 – 292.
References Britt, D.W., Hlady, V., 1997. In situ atomic force microscope imag-
ing of calcite etch morphology change in undersaturated and 1-
Akin, G.W., Lagerwerff, J.V., 1965. Calcium carbonate equilibrium hydroxyethylidene-1, 1-diphosphonic acid poisoned solutions.
in aqueous solutions open to the air: II. Enhanced solubility of a Langmuir 13, 1873 – 1876.
CaCO3 in the presence of Mg2 + and SO2 4 . Geochim. Cos- Brunauer, S., Emmett, P.H., Teller, E., 1938. Absorption of gases
mochim. Acta 29, 353 – 360. and multimolecular layers. Journal of the American Chemical
Arakaki, T., Mucci, A., 1995. A continuous and mechanistic repre- Society 60, 309 – 319.
sentation of calcite reaction-controlled kinetics in dilute solu- Buhmann, D., Dreybrodt, W., 1987. Calcite dissolution kinetics in
tions at 25 jC and 1 atm total pressure. Aquat. Geochem. 1, the system H2O – CO 2 CaCO3 with participation of foreign
105 – 130. ions. Chem. Geol. 64, 89 – 102.
J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84 81
Burton, W.K., Cabrera, N., Frank, F.C., 1951. The growth of crys- for two-component solids reacting with fixed composition in an
tals and the equilibrium structure of their surfaces. Philos. Trans. aqueous phase-example: the magnesian calcites. Am. J. Sci.
R. Soc. London, Ser. A 243, 299 – 358. 278, 1469 – 1474.
Busenberg, E., Plummer, L.N., 1982. The kinetics of dissolution of Garrels, R.M., Thompson, M.E., Siever, R., 1960. Stability of some
dolomite in CO2 – H2O Systems at 1.5 to 65 jC and 0 to 1 atm carbonates at 25 jC and one atmosphere total pressure. Am. J.
PCO2. Am. J. Sci. 282, 45 – 78. Sci. 258, 402 – 418.
Busenberg, E., Plummer, L.N., 1986. A comparative study of the Garrels, R.M., Thompson, M.E., Siever, R., 1961. Control of car-
dissolution and precipitation kinetics of calcite and aragonite. In: bonate solubility by carbonate complexes. Am. J. Sci. 259,
Mumton, F.A. (Ed.), Studies in Diagenesis. U. S. Geol. Surv. 24 – 25.
Bull., pp. 139 – 168. Gautelier, M., Oelkers, E.H., Schott, J., 1999. An experimental-
Chai, L., Navrotsky, A., Reeder, R.J., 1995. Energetics of calcium- study of dolomite dissolution rates as a function of pH from
rich dolomite. Geochim. Cosmochim. Acta 59, 939 – 944. 0.5 to 5 and temperature from 25 to 80 jC. Chem. Geol. 157,
Chou, L., Garrels, R.M., Wollast, R., 1989. Comparative study of the 13 – 26.
kinetics and mechanisms of dissolution of carbonate minerals. Gautier, J.-M., Oelkers, E.H., Schott, J., 2001. Are quartz dissolu-
Chem. Geol. 78, 269 – 282. tion rates proportional to B.E.T. surface areas? Geochemica et
Compton, R.G., Sanders, G.H.W., 1993. The dissolution of calcite Cosmochimica Acta 65, 1059 – 1070.
in aqueous acid: the influence of humic species. J. Colloid Inter- Gratz, A.J., Manne, S., Hansma, P.K., 1991. Atomic force micro-
face Sci. 158, 439 – 445. scopy of atomic-scale ledges and etch pits formed during dis-
Compton, R.G., Daly, P.J., House, W.A., 1986. The dissolution of solution of quartz. Science 251, 1343 – 1346.
Iceland spar crystals: the effect of surface morphology. J. Col- Gratz, A.J., Hillner, P.E., Hasma, P.K., 1993. Step dynamics and spiral
loid Interface Sci. 113, 12 – 20. growth on calcite. Geochim. Cosmochim. Acta 57, 491 – 495.
Compton, R.G., Pritchard, K.L., Unwin, P.R., 1989. The dissolution Gross, M.G., 1964. Variations in the O18/O16 and C13/C12 ratios of
of calcite in acid waters: mass transport versus surface control. diagenetically altered limestones in the Bermuda Islands. J. Geol.
Freshwater Biol. 22, 286 – 288. 72, 170 – 194.
Davis, K.J., Dove, P.M., De Yoreo, J.J., 2000. The role of Mg2 + as Gutjahr, A., Dabringhaus, H., Lacmann, R., 1996a. Studies of the
an impurity in calcite growth. Science 290, 1134 – 1137. growth and dissolution kinetics of the CaCO3 polymorphs calcite
Dove, P.M., Hochella, M.F., 1993. Calcite precipitation mecha- and aragonite: I. Growth and dissolution rates in water. J. Cryst.
nisms and inhibition by orthophosphate: in situ observations Growth 158, 296 – 309.
by Scanning Force Microscopy. Geochim. Cosmochim. Acta Gutjahr, A., Dabringhaus, H., Lacmann, R., 1996b. Studies of the
57, 705 – 714. growth and dissolution kinetics of the CaCO3 polymorphs
Dove, P.M., Platt, F.M., 1996. Compatible real-time reaction rates calcite and aragonite: II. The influence of divalent cation ad-
for in situ imaging of mineral – water interactions using scanning ditives on the growth and dissolution rates. J. Cryst. Growth
force microscopy. Chem. Geol. 127, 331 – 338. 158, 310 – 315.
Drake, B., Hellmann, R., 1991. Atomic force microscopy imaging Hales, B., Emerson, S., 1997. Evidence in support of first-order
of the albite (010) surface. Am. Mineral. 76, 1773 – 1776. dissolution kinetics of calcite in seawater. Earth Planet. Sci. Lett.
Dreybrodt, W., Buhmann, D., 1991. A mass transfer model for 148, 317 – 327.
dissolution and precipitation of calcite from solutions in turbu- Helgeson, H.C., Murphy, W.M., Aagard, P., 1984. Thermodynamic
lent motion. Chem. Geol. 90, 107 – 122. and kinetic constraints on reaction rates among minerals and
Dreybrodt, W., Lauckner, J., Zaihua, L., Svensson, U., Buhmann, aqueous solutions. Rate constants, effective surface area, and
D., 1996. The kinetics of the reaction CO2 + H2O (goes to) the hydrolysis of feldspar. Geochim. Cosmochim. Acta 48,
H + + HCO3 as one of the rate limiting steps for the dissolution 2405 – 2432.
of calcite in the system H2O – CO2 – CaCO3. Geochim. Cosmo- Hillner, P.E., Manne, S., Gratz, A.J., Hasma, P.K., 1992a. AFM
chim. Acta 60, 3375 – 3381. images of dissolution and growth on a calcite crystal. Ultrami-
Edmond, J.M., 1971. An interpretation of calcite spheres experi- croscopy 42 – 44, 1387 – 1393.
ment. EOS, Trans. Am. Geophys. 52, 256. Hillner, P.E., Gratz, A.J., Manne, S., Hasma, P.K., 1992b. Atomic-
Erga, O., Terjesen, S.G., 1956. Kinetics of heterogeneous reaction scale imaging of calcite growth and dissolution in real time.
of calcium bicarbonate formation, with specidal reference to Geology 20, 359 – 362.
copper ion inhibition. Acta Chem. Scand. 10, 872 – 874. Hochella Jr., M.F., Eggleston, C.M., Elings, V.B., Thompson, M.S.,
Friedman, G.M., 1964. Early diagenesis and lithification in carbo- 1990. Atomic structure and morphology of the albite {010}
nate sediments. J. Sediment. Petrol. 34, 777 – 813. surface: an atomic-force microscope and electron diffraction
Friedman, G.M., Sanders, J.E., 1967. Origin and occurrence of study. Am. Mineral. 75, 723 – 730.
dolostones. In: Chilingar, G.V., Bissel, H.J., Fairbridge, R.W. Honjo, S., Erez, J., 1978. Dissolution rates of calcium carbonate in
(Eds.), Carbonate Rocks: Origin, Occurrence and Classification. the deep ocean: an in-situ experiment in the North Atlantic.
Elsevier, Amsterdam, pp. 267 – 348. Earth Planet. Sci. Lett. 40, 287 – 300.
Gaffey, S.J., 1988. Water in skeletal carbonates. J. Sediment. Petrol. Johnson, J.W., Oelkers, E.H., Helgeson, H.C., 1992. SUPCRT92: a
58, 397 – 414. software package for calculating the standard molal thermody-
Garrels, R.M., Wollast, R., 1978. Discussion of equilibrium criteria namic properties of minerals, gases, aqueous species, and reac-
82 J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84
tions from 1 to 5000 bar and 0 to 1000 jC. Comput. Geosci. 18, Lüttge, A., Winkler, U., Lasaga, A.C., 2000. An interferometric
899 – 947. study of the dissolution kinetics of rhombohedral faces in dolo-
Johnsson, P.A., Eggleston, C.M., Hochella, M.F., 1991. Imaging mite. Geochim. Cosmochim. Acta, in review.
molecular-scale structure and microtopography of hematite with Lüttge, A., Winkler, U., Lasaga, A.C., 2002. An interferometric
the atomic force microscope. Am. Mineral. 76, 1442 – 1445. study of the dolomite dissolution: A new conceptual model for
Jordan, G., Rammensee, W., 1998. Dissolution rates of calcite mineral dissolution. Geochimica et Cosmochimica Acta, in press.
(101̄4) obtained by scanning force microscopy; microtopogra- MacInnis, I.N., Brantley, S.L., 1992. The role of dislocations and
phy-based dissolution kinetics on surfaces with anisotropic step surface morphology in calcite dissolution. Geochim. Cosmo-
velocities. Geochim. Cosmochim. Acta 62, 941 – 947. chim. Acta 56, 1113 – 1126.
Kier, R.S., 1980. The dissolution kinetics of biogenic calcium car- MacInnis, I.N., Brantley, S.L., 1993. Development of etch pit size
bonates in seawater. Geochim. Cosmochim. Acta 44, 241 – 252. distributions on dissolving minerals. Chem. Geol. 105, 31 – 49.
Kinsman, D.J.J., 1965. Dolomitization and evaporite development, Mackenzie, F.T., Bischoff, W.D., Bishop, F.C., Loijens, M., Schoon-
including anhydrite, in lagoonal sediments, Persian Gulf. Geol. maker, J., Wollast, R., 1983. Magnesian calcites; low-temper-
Soc. Am. Spec. Pap. 82, 108 – 109. ature occurrence, solubility and solid – solution behavior. In:
Kushnir, J., 1983. The rate of dissolution of carbonate minerals in Reeder, R.J. (Ed.), Carbonates: Mineralogy and Chemistry.
the presence of sulfate ions. Annual Meeting of the Geological Rev. Mineral. Mineralogical Society of America, Washington,
Society of America, Abstract, 620. DC, vol. 11, pp. 97 – 144.
Lafon, G.M., 1978. Equilibrium criteria for two-component solids Marti, O., Drake, B., Hansma, P.K., 1987. Atomic force microscopy
reacting with fixed composition in an aqueous phase; example, of liquid-covered surfaces: atomic resolution images. Appl.
the magnesian calcites; discussion. Am. J. Sci. 278, 1455 – 1468. Phys. Lett. 51, 484 – 486.
Land, L.S., 1966. Diagenesis of metastable carbonates. Thesis, Milliman, J.D., 1977. Dissolution of calcium carbonate in the Sar-
Lehigh Univ., Bethlehem, PA. gasso Sea (Northwest Atlantic). In: Andersen, N.R., Malahoff,
Land, L.S., 1967. Diagenesis of skeletal carbonates. J. Sediment. A. (Eds.), The Fate of Fossil Fuel CO2 in the Oceans. Plenum,
Petrol. 37, 914 – 930. New York, NY, pp. 641 – 654.
Land, L.S., Epstein, S., 1970. Late Pleistocene diagenesis and Möller, P., 1973. Determination of the composition of surface layers
dolomitization, north Jamaica. Sedimentology 14, 187 – 200. of calcite in solutions containing Mg2 +. J. Inorg. Nuclear
Lasaga, A.C., 1998. Kinetic Theory in the Earth Sciences. Princeton Chem. 35, 395 – 401.
Univ. Press, Princeton, NJ. Möller, P., Rajagopalan, G., 1975. Precipitation kinetics of CaCO3
Lasaga, A.C., Luttge, A., 2001. Variation of crystal dissolution rate in presence of Mg2 + ions. Z. Phys. Chem., Neue Folge 94,
based on a dissolution stepwave model. Science 291, 2400 – 2404. 297 – 314.
Lea, A.S., Amonette, J.E., Baer, D.R., Liang, Y., Colton, N.G., Moore, C.H., 1989. Carbonate Diagenesis and Porosity. Develop-
2001. Microscopic effects of carbonate, manganese, and stron- ments in Sedimentology 46, Elsevier, New York, NY.
tium ions on calcite dissolution. Geochim. Cosmochim. Acta 65, Morse J.W., 1973. The dissolution kinetics of calcite: a kinetic origin
369 – 379. for the lysocline. PhD Thesis, Yale University, New Haven, CT.
Liang, Y., Baer, D.R., 1997. Anisotropic dissolution at the CaCO3 Morse, J.W., 1974. Dissolution kinetics of calcium carbonate in sea
(101̄4) – water interface. Surf. Sci. 373, 275 – 287. water: V. Effects of natural inhibitors and the position of the
Liang, Y., Baer, D.R., McCoy, J.M., Amonette, J.E., LaFemina, J.P., chemical lysocline. Am. J. Sci. 274, 638 – 647.
1996a. Dissolution kinetics at the calcite – water interface. Geo- Morse, J.W., 1978. Dissolution kinetics of calcium carbonate in sea-
chim. Cosmochim. Acta 60, 4883 – 4887. water: IV. The near-equilibrium dissolution kinetics of calcium
Liang, Y., Baer, D.R., McCoy, J.M., LaFemina, J.P., 1996b. Inter- carbonate-rich deep sea sediments. Am. J. Sci. 278, 344 – 353.
play between step velocity and morphology during the dissolu- Morse, J.W., 1983. The kinetics of calcium carbonate dissolution
tion of CaCO3 surface. J. Vacuum Sci. Technol. 14, 1363 – 1375. and precipitation. In: Reeder, R.J. (Ed.), Carbonates: Mineral-
Lindgreen, H., Gaenees, J., Hansen, P.L., Besenbacker, F., Laes- ogy and Chemistry Rev. Mineral. Mineralogical Society of
gaard, E., Stensgaard, I., Gould, S.A.C., Hansma, P.K., 1991. America, Washington, DC, vol. 11, pp. 227 – 264.
Ultrafine particles of North Sea illite/smectite clay minerals in- Morse, J.W., Berner, R.A., 1972. Dissolution kinetics of calcium
vestigated by STM and AFM. Am. Mineral. 76, 1218 – 1222. carbonate in seawater: II. A kinetic origin for lysocline. Am. J.
Liu, Z., Dreybrodt, W., 1997. Dissolution kinetics of calcium car- Sci. 272, 840 – 851.
bonate minerals in H2O – CO2 solutions in turbulent flow: the Morse, J.W., Berner, R.A., 1979. The chemistry of calcium carbo-
role of the diffusion boundary layer and the slow reaction nate in the deep oceans. In: Jenne, E. (Ed.), Chemical Model-
H2O + CO2 X H + + HCO3 . Geochim. Cosmochim. Acta 61, ing — Speciation, Sorption, Solubility, and Kinetics in Aqueous
2879 – 2889. Systems. American Chemical Society, Washington, DC, pp.
Lund, K., Fogler, H.S., McCune, C.C., 1973. Acidization — I. The 499 – 535.
dissolution of dolomite in hydrochloric acid. Chem. Eng. Sci. Morse, J.W., Mackenzie, F.T., 1990. Geochemistry of Sedimentary
28, 691 – 700. Carbonates. Elsevier, Amsterdam.
Lüttge, A., Bolton, E.W., Lasaga, A.C., 1999. An interoferometric Morse, J.W., deKanel, J., Harris, J., 1979. Dissolution kinetics of
study of the dissolution kinetics of anorthite: the role of reactive calcium carbonate in seawater: VII. The dissolution kinetics of
surface area. Am. J. Sci. 299, 652 – 678. synthetic aragonite and pteropod tests. Am. J. Sci. 279, 482 – 502.
J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84 83
Mucci, A., Morse, J.W., 1984. The solubility of calcite in seawater anions: I. Inhibition by phosphate and glycerophophate ions at
solutions of various magnesium concentration, Iz = 0.697 m at pH 8.8 and 25 jC. J. Cryst. Growth 41, 287 – 295.
25 degrees C and one atmosphere total pressure. Geochim. Reddy, M.M., Wang, K.K., 1980. Crystallization of calcium carbo-
Cosmochim. Acta 48, 815 – 822. nate in the presence of metal ions: I. Inhibition by magnesium
Mucci, A., Canuel, R., Zhong, S., 1989. The solubility of calcite and ion at pH 8.8 and 25 jC. J. Cryst. Growth 50, 470 – 480.
aragonite in sulfate-free seawater and the seeded growth kinetics Rickard, D.T., Sjöberg, E.L., 1983. Mixed kinetic control of calcite
and composition of precipitates at 25 jC. Geochim. Cosmochim. dissolution rates. Am. J. Sci. 283, 815 – 830.
Acta 74, 309 – 329. Robie, R.A., Hemingway, B.S., Fisher, J.S., 1979. Thermodynamic
Murray, J., Renard, A.F., 1891. Report of the deep-sea deposits. properties of minerals and related substances at 298.15 K and 1
From the on the Scientific Results of the Voyage H.M.S. Chal- Bar (105 Pascals) pressure and at higher temperatures. U.S. Geol.
lenger, Great Britain, Challenger Office. Surv. Bull. 1452, 1 – 456.
Nancollas, G.H., Kazmierczak, T.F., Schuttringer, E., 1981. A con- Rufe, E., Hochella, M.F., 1999. Quantitative assessment of reactive
trolled composition study of calcium carbonate crystal growth: surface area of phlogopite during acid dissolution. Science 285,
the influence of scale inhibitors. Corrosion 37, 76 – 81. 874 – 876.
Navrotsky, A., Capobianco, C., 1987. Enthalpies of formation of Salem, M.R., Mangood, A.H., Hamdona, S.K., 1994. Dissolution of
dolomite and of magnesian calcites. Am. Mineral. 72, 782 – 787. calcite crystals in the presence of some metal ions. J. Mater. Sci.
Nestaas, I., Terjesen, S.G., 1969. The inhibiting effect of scandium 29, 6463 – 6467.
ions upon the dissolution of calcium carbonate. Acta Chem. Schmalz, R.F., 1967. Kinetics and diagenesis of carbonate sedi-
Scand. 23, 2519 – 2531. ments. J. Sediment. Petrol. 37, 60 – 67.
Nielson, A.E., 1964. Kinetics of Precipitation. Pergamon, NY. Schott, J., Brantley, S., Drear, D., Guy, C., Borcsik, M., Willaime,
Paquette, J., Reeder, R.J., 1990. Single-crystal X-ray structure re- C., 1989. Dissolution kinetics of strained calcite. Geochim. Cos-
finements of two biogenic magnesian calcite crystals. Am. Min- mochim. Acta 53, 373 – 382.
eral. 75, 1151 – 1158. Schroeder, J.H., Siegel, F.R., 1969. Experimental dissolution of cal-
Park, N.S., Kim, M.W., Langford, S.C., Dickinson, J.T., 1996. Tri- cium, magnesium and strontium from Recent biogenic carbo-
bological enhancement of CaCO3 dissolution during scanning nates: a model for diagenesis. Bull. Am. Assoc. Pet. Geol. 53, 741.
force microscopy. Langmuir 12, 4599 – 4604. Shiraki, R., Rock, P.A., Casey, W.H., 2000. Dissolution kinetics of
Peterson, M.N.A., 1966. Calcite: rates of dissolution in a vertical calcite in 0.1 M NaCl solution at room temperature: an atomic
profile in the central Pacific. Science 154, 1542 – 1544. force microscopic (AFM) study. Aquat. Geochem. 6, 87 – 108.
Plath, D.C., Johnson, K.S., Pytkowicz, R.M., 1980. The solubility Sjöberg, E.L., 1976. A fundamental equation for calcite dissolution
of calcite probably containing magnesium in seawater. Mar. kinetics. Geochim. Cosmochim. Acta 40, 441 – 447.
Chem. 10, 9 – 29. Sjöberg, E.L., 1978. Kinetics and mechanism of calcite dissolution
Plummer, L.N., Mackenzie, F.T., 1974. Predicting mineral solubility in aqueous solutions at low temperatures. Stockholm Contrib.
from rate data — application to the dissolution of magnesium Geol. 32, 92 pp.
calcites. Am. J. Sci. 274, 61 – 83. Sjöberg, E.L., Rickard, D.T., 1983. The influence of experimental
Plummer, L.N., Wigley, T.M.L., 1976. The dissolution of calcite in design on the rate of calcite dissolution. Geochim. Cosmochim.
CO2-saturated solutions at 25 jC and 1 atmosphere total pres- Acta 47, 2281 – 2286.
sure. Geochim. Cosmochim. Acta 40, 191 – 201. Sjöberg, E.L., Rickard, D.T., 1984a. Calcite dissolution kinetics:
Plummer, L.N., Wigley, T.M.L., Parkhurst, D.L., 1978. The kinetics surface specification and the origin of the variable pH depend-
of calcite dissolution in CO2 – water systems at 5j and 60 jC ence. Chem. Geol. 42, 119 – 136.
and 0.0 to 1.0 ATM CO2. Am. J. Sci. 278, 179 – 216. Sjöberg, E.L., Rickard, D.T., 1984b. Temperature dependence of
Plummer, L.N., Parkhurst, D.L., Wigley, T.M.L., 1979. Critical calcite dissolution kinetics between 1 and 62 jC at pH 2.7
review of the kinetics of calcite dissolution and precipitation. to 8.4 in aqueous solutions. Geochim. Cosmochim. Acta 48,
In: Jenne, E. (Ed.), Chemical Modeling — Speciation, Sorption, 485 – 493.
Solubility and Kinetics in Aqueous Systems. American Chem- Sjöberg, E.L., Rickard, D.T., 1985. The effect of added dissolved
ical Society, Washington, DC, pp. 537 – 573. calcium on calcite dissolution kinetics in aqueous solutions at
Pokrovsky, O.S., Schott, J., 1999. Processes at the magnesium-bear- 25 jC. Chem. Geol. 49, 405 – 413.
ing carbonates/solution interface: II. Kinetics and mechanism of Suess, E., 1970. Interaction of organic compounds with calcium
magnesite dissolution. Geochim. Cosmochim. Acta 63, 881 – 897. carbonate: I. Association phenomena and geochemical implica-
Pokrovsky, O.S., Schott, J., Fabien, T., 1999a. Dolomite surface tions. Geochim. Cosmochim. Acta 34, 157 – 168.
speciation and reactivity in aquatic systems. Geochim. Cosmo- Svensson, U., Dreybrodt, W., 1992. Dissolution kinetics of natural
chim. Acta 63, 3133 – 3143. calcite minerals in CO2 – water systems approaching calcite
Pokrovsky, O.S., Schott, J., Fabien, T., 1999b. Processes at the equilibrium. Chem. Geol. 100, 129 – 145.
magnesium-bearing carbonates/solution interface: I. A surface Talman, S.J., Gunter, W.D., 1992. Rates of dolomite dissolution in
speciation model for magnesite. Geochim. Cosmochim. Acta CO2 and HCl bearing solutions from 100 – 200 degrees C. In:
63, 863 – 880. Kharaka, Y.K., Maest, A.S. (Eds.), Proceedings of the 7th Inter-
Reddy, M.M., 1977. Crystallization of calcium carbonate in the national Symposium on Water – Rock Interaction, Volume 1,
presence of trace concentrations of phosphorus containing Low Temperature Environments. International Association of
84 J.W. Morse, R.S. Arvidson / Earth-Science Reviews 58 (2002) 51–84
Geochemistry and Cosmochemistry and Alberta Research carbonates during dissolution: effect of grain size. J. Sediment.
Council, Sub-Group on Water – Rock Interaction, Edmonton, Petrol. 54, 1081 – 1090.
AB, 119 – 122. Walter, L.M., Morse, J.W., 1985. The dissolution kinetics of shallow
Terjesen, S.G., Erga, O., Thorsen, G., Ve, A., 1961. II. Phase boun- marine carbonates in seawater: a laboratory study. Geochim.
dary processes as rate determining steps in the reaction between Cosmochim. Acta 49, 1503 – 1513.
solids and liquids. The inhibitory action of metal ions on the Weisenhorn, A.L., MacDougall, J.E., Gould, S.A.C., Cox, S.D.,
formation calcium bicarbonate by the reaction of calcite with Wise, W.S., Massie, J., Maivald, P., Elings, V.B., Stucky, G.D.,
aqueous carbon dioxide. Chem. Eng. Sci. 74, 277 – 288. Hansma, P.K., 1990. Imaging and manipulating molecules on a
Thorstenson, D.C., Plummer, L.N., 1977. Equilibrium criteria for zeolite surface with an atomic force microscope. Science 247,
two-component solids reacting with fixed composition in an 1330 – 1333.
aqueous phase; example, the magnesian calcites. Am. J. Sci. Weyl, P.K., 1958. The solution kinetics of calcite. J. Geol. 66,
277, 1203 – 1223. 163 – 176.
Thorstenson, D.C., Plummer, L.N., 1978. Equilibrium criteria for Weyl, P.K., 1965. The solution kinetics of calcite. ERP Publ. 428,
two-component solids reacting with fixed composition in an Shell Development Company , 1 – 59.
aqueous phase; example, the magnesian calcites; reply. Am. J. White, A.F., Peterson, M.L., 1990. Role of reactive-surface-area char-
Sci. 278, 1478 – 1488. acterization in geochemical kinetic models. In: Melchior, D.C.,
Van Cappellen, P., Charlet, L., Stumm, W., Wersin, P., 1993. A Bassett, R.L. (Eds.), Chemical Modeling of Aqueous Systems II
surface complexation model of the carbonate mineral – aqueous Am. Chem. Soc., Symp. Ser. American Chemical Society, Wash-
solution interface. Geochim. Cosmochim. Acta 57, 3505 – 3518. ington, DC, vol. 416, pp. 461 – 475.
Van Name, R.G., Hill, D.U., 1916. On the rates of solutions of Wollast, R., Reinhard-Derie, D., 1977. Equilibrium and mechanism
metals in ferric salts and in chromic acid. Am. J. Sci. 42, of dissolution of Mg-calcites. In: Andersen, N.R., Malahoff, A.
301 – 332. (Eds.), The Fate of Fossil Fuel CO2 in the Oceans. Plenum, New
Walter, L.M., Hanor, J.S., 1979. Effect of orthophosphate on the York, NY, pp. 479 – 493.
dissolution kinetics of biogenic magnesian calcite. Geochim. Zhang, Y., Grattoni, C.A., 1998. Comment of ‘‘Precipitation ki-
Cosmochim. Acta 43, 1377 – 1385. netics of calcite in the system CaCO3 – H2O – CO2: The conver-
Walter, L.M., Morse, J.W., 1984a. Magnesium calcite stabilities: a sion to CO2 by the slow process H + + HCO3 ! CO2 + H2O as
reevaluation. Geochim. Cosmochim. Acta 48, 1059 – 1069. a rate limiting step’’ by W. Dreybrodt, L. Eisenlohr, B. Madry,
Walter, L.M., Morse, J.W., 1984b. Reactive surface area of skeletal and S. Ringer. Geochim. Cosmochim. Acta 62, 3789 – 3790.