You are on page 1of 11

American Mineralogist, Volume 87, pages 679–689, 2002

Surface structures, stabilities, and growth of magnesian calcites: A computational


investigation from the perspective of dolomite formation

NORA H. DE LEEUW*

Department of Chemistry, University of Reading, Whiteknights, Reading RG6 6AD, U.K.

ABSTRACT
Atomistic computer modeling methods are used to investigate the surface structures and stabili-
ties of magnesite MgCO3 and dolomite MgCa(CO3)2 in comparison with calcite, CaCO3. The sur-
faces are generally more stable than magnesite, but less so than calcite. Upon hydration, the magnesite

{1014} surface becomes more stable than either calcite or dolomite, due to the less favorable inter-
actions between dolomite surface ions and adsorbing water molecules. The high affinity of the sur-
face magnesium ions for adsorbing water molecules, shown by the large average hydration energy
over all surfaces for magnesite (134 kJ/mol), is negated in the case of dolomite by the inaccessibility
of the surface magnesium ions, due to surface relaxations and rotation of the carbonate groups. In
addition, the larger lattice spacing of dolomite compared to magnesite disrupts the extensive net-
work of hydrogen bonding between the adsorbed water molecules that are present on the latter,
leading to similar average hydration energies for calcite, dolomite, and magnesian calcite surfaces
(80–101 kJ/mol). Not only are the dolomite planes under aqueous conditions less stable than their
calcite counterparts, dolomite-like surfaces of disordered magnesian calcites are calculated to be
more stable than either pure calcite or dolomite surfaces. Extensive molecular dynamics simulations
of the growth of CaCO3 and MgCO3 at two experimentally observed calcite steps show that, on
thermodynamic grounds, Mg is easily incorporated into a growing calcite crystal, but will then
inhibit incorporation of further Ca. Furthermore, although the calculations suggest a small energetic
advantage (~10 kJ/mol) to ordering of the Ca and Mg ions on the obtuse growth step, a disordered
arrangement is favored at the acute growth step and incorporation of Mg is always energetically
more favorable than Ca, in accord with the occurrence of highly magnesian calcites in nature.

INTRODUCTION such as hot springs, in sediments of salt lakes, and in muds


Dolomite, an ordered Ca-Mg carbonate mineral, is found from salt lagoons undergoing strong solar evaporation, sedi-
extensively in ancient rocks in the Earth’s crust (e.g., Deer et mentary dolomite does not form in the normal present-day
al. 1992) and calcite and dolomite together account for more marine environment and it is notoriously difficult to crystal-
than 90% of the natural carbonates in rocks (Reeder 1990). lize under laboratory conditions (Krauskopf and Bird 1995).
Despite several important occurrences in metamorphic rocks Laboratory precipitation has been achieved at high tempera-
(e.g., Letargo et al. 1995; Ferry 1996, 2000, 2001), dolomite is tures (e.g., Katz and Matthew 1977; Baker and Kastner 1981;
typically a sedimentary mineral, yet the mechanism of its growth Sibley et al. 1994), at extreme supersaturations (e.g.,
under sedimentary conditions is unknown, apart from the fact Liebermann 1967; Nordeng and Sibley 1994) or from solu-
that there is no geologic evidence indicating that its formation tions with pH values greater than 9.5 and containing high con-
took place under unusual conditions of temperature or pres- centrations of SO2– –
4 and NO3 (Sibley et al. 1994; Krauskopf

sure (Krauskopf and Bird 1995). Some primary dolomites are and Bird 1995), but problems persist when attempting to crys-
formed at the outset as evaporite deposits, but most dolomitic tallize dolomite under the usual mild conditions applicable to
sediments are produced from the reaction of calcium carbon- the most common carbonate mineral calcite (Brady et al. 1996).
ate deposits (e.g., isomorphous calcite) with magnesium solu- It is these extreme conditions represented by both the labora-
tions. These secondary dolomites are either formed early on tory experiments and field occurrences, compared with usual
from unconsolidated calcite sediments on the sea floor, or much sedimentary environments, which makes the abundance of do-
later when magnesian solutions have entered through faults and lomite in sedimentary rocks so mysterious.
joints in the calcium carbonate rock (e.g., Wenk 1993). How- Not surprisingly, dolomite and its origins has been the sub-
ever, although formation of dolomite as a primary precipitate ject of numerous investigations, both experimentally and theo-
in nature has been reported in several unusual environments, retically. For example, dissolution experiments by Bertram et
al. (1991) showed that the solubility of synthetic high magne-
* E-mail: n.h.deleeuw@reading.ac.uk sian calcite decreases with increasing temperature, but that re-
0003-004X/02/0506–679$05.00 679
680 DE LEEUW: MODELING DOLOMITE AND MAGNESIAN CALCITES

crystallization at low temperatures into dolomite and low-mag- lytical functions. The electronic polarizability of the ions is
nesian calcite should be thermodynamically favored, while included via the shell model of Dick and Overhauser (1958) in
experimental work on nucleation and growth kinetics of cal- which each polarizable ion, in this case the O atom ion, is rep-
cium and magnesium carbonates by Nordeng and Sibley (1994) resented by a core and a massless shell, connected by a spring.
showed that the surface free energy of high-magnesium calcite The polarizability of the model ion is then determined by the
is less than that of dolomite. Investigations of structural as- spring constant and the charges of the core and shell. When
pects of dolomite include Ross and Reeder’s (1992) X-ray in- necessary, angle-dependent forces are included to allow direc-
tensity data refining the structural parameters of stoichiometric tionality of bonding as, for example, in the model of the cova-
and ferroan dolomites and high pressure X-ray diffraction stud- lent carbonate anion developed by Pavese et al. (1996). For the
ies by Fiquet et al. (1994) to obtain bulk moduli for dolomite calcium carbonate materials, the present model uses the short-
and other carbonates, while Williams et al. (1992) measured range interactions derived empirically by Pavese et al. (1996)
vibrational spectra of magnesite and dolomite. More recent stud- in their study of the thermal dependence of structural and elas-
ies of dolomites have investigated surface growth processes, tic properties of calcite. Although this potential was fitted to
for example the transmission electron microscopy (TEM) stud- bulk properties, it is generally possible for ionic materials to
ies by Paquette et al. (1999) of growth hillocks and surface transfer potential parameters to surface calculations. In semi-
microtopography, and scanning microscopy investigations of conductors, where the surface involves broken bonds, and met-
etch pits, layers, and islands on synthetic and natural dolomite als, where the surface means a sudden change in the electron
surfaces by Kessels et al. (2000). density, there is often a problem with transferability from bulk
Theoretical models based on experimental observations in- potential parameters to surfaces. However, in ionic materials
clude the surface complexation model by Brady et al. (1996), after relaxation, the Madelung potentials are 90% or more of
who proposed the formation of Mg-sulfate complexes as a pos- the bulk values and hence the change in ionic radii is negli-
sible path to dolomite growth and surface speciation models gible. Oliver et al. (1996), for example, used bulk derived po-
for both magnesite and dolomite based on measurements of tentials to model WO3 surfaces and found excellent agreement
surface charge as a function of pH (Pokrovsky et al. 1999a, between calculated and experimental surface structures, ob-
1999b). Previous computer modeling studies have addressed tained by scanning tunneling microscopy. Because the surfaces
the surface structures and stabilities of dehydrated dolomite considered in this work leave the carbonate groups intact, the
relative to calcite (Titiloye et al. 1998) while Wright et al. (2001) bulk derived potential model will be adequate. Indeed, the po-

compared the structure of the planar (1014) surface to those of tential model derived for calcite by Pavese et al. (1996) has
magnesite and calcite. Fisler et al. (2000) concentrated on the been used successfully in numerous surface simulations, not
ordering of Mg and Ca in the bulk mineral, following up ear- only of calcite, but also of the less ubiquitous calcium carbon-
lier experimental work on the behavior of Mg diffusion in bulk ate phases aragonite and vaterite (e.g., de Leeuw and Parker
calcite (Fisler and Cygan 1999). 1997, 1998a; Titiloye et al. 1998). The experimental surface
In this article I use a combination of classical simulation energies and structural features were reproduced accurately (de
techniques, employing established potential models (Pavese et Leeuw and Parker 1997), while the calculated morphologies
al. 1996; de Leeuw and Parker 1998a, 2000). I investigate the of the three calcium carbonate phases, which are dependent on
surface structures and stabilities of a range of dry and hydrated the relative stabilities of the various surfaces, agreed well with
dolomite surfaces, which are observed in the experimental mor- experimentally determined crystal morphologies (de Leeuw and
phology (Deer et al. 1992). I compare the same with their coun- Parker 1998a). The parameters used for the Mg-CO3 interac-
terparts in the pure end-members calcite and magnesite and a tions and those between Mg and water were derived in an ear-
disordered magnesian calcite. In addition, as dolomite forma- lier work (de Leeuw and Parker 2000), using a combination of
tion often occurs from the interaction of calcium carbonate with static lattice energy minimization calculations of the magnes-
Mg solutions (Krauskopf and Bird 1995), I also calculate the ite (MgCO3) crystal and molecular dynamics (MD) simulations
energetics of incorporating Mg at two experimentally observed of the Mg2+ ion in liquid water. The intra- and intermolecular

growth steps on the calcite {1014} surface under aqueous con- interactions of the water molecules were derived using MD
ditions, taking into account the ordering required to form the simulations of liquid water, the details of which are described
dolomite crystal. My calculations provide insight into the ef- in a previous paper on modeling MgO surfaces in liquid water
fect of the different surface ions on the relative surface stabili- (de Leeuw and Parker 1998b). The interactions between water
ties of the carbonate phases, which has implications for the and calcite surfaces were first derived in a previous work on
mineral morphologies, and the relative enthalpies of growth of water adsorption at calcite surfaces (de Leeuw and Parker 1997).
ordered and disordered magnesian calcites. The full potential model is given in Table 1.
The simulation methods used in this work are a combina-
THEORETICAL METHODS tion of static lattice energy minimization techniques and MD
Atomistic simulation techniques are based on the Born simulations. The energy minimization codes employed for the
model of solids (Born and Huang 1954), which assumes that calculation of the bulk crystal properties were PARAPOCS
the ions in the crystal interact via long-range electrostatic forces (Parker and Price 1989) and GULP (Gale 1997), while the sur-
and short-range forces, including both repulsions and the van faces were modeled using the code METADISE (Watson et al.
der Waals attractions between neighboring electron charge 1996), which is designed to model dislocations, interfaces, and
clouds. These short-range forces are described by simple ana- surfaces. Following the approach of Tasker (1979), the crystal
DE LEEUW: MODELING DOLOMITE AND MAGNESIAN CALCITES 681

consists of a series of charged planes parallel to the surface test runs with various masses for the O atom shell (from 0.05
and periodic in two dimensions. The crystal is divided into two to 0.5 au), 0.2 au was chosen which is small compared to the
blocks, each comprising two regions, region I and region II. mass of the hydrogen atom of 1.0 au, ensuring that there would
Region I contains those atoms near the surface, which are al- be no exchange of energy between vibrations of O atom core
lowed to relax to their mechanical equilibrium, while region II and shell with O atom and hydrogen vibrations (de Leeuw and
contains those atoms further away, which represent the rest of Parker 1998b). However, due to this small shell mass the MD
the crystal and are kept fixed at their bulk equilibrium posi- simulations needed to be run with a small timestep of 0.2 fs to
tion. Inclusion of region II is necessary to ensure that the po- keep the system stable. The stepped calcite surfaces were mod-
tential of an ion at the bottom of region I is modeled correctly. eled as a repeating slab, containing 120 CaCO3 units and a gap,
The DL_POLY code (Forester and Smith 1995) was used containing 48 water molecules, giving a total simulation cell

for the MD simulations of growth steps on the calcite {1014} of 1152 species. MD calculations of calcite dissolution (de
surface, where the integration algorithms are based around the Leeuw et al. 1999) showed that the size of these simulation
Verlet leap-frog scheme (Verlet 1967). The Nosé-Hoover algo- cells (where the steps are more than 10 Å apart) ensured that
rithm (Nosé 1984; Hoover 1985) was used for the thermostat there are no interactions between the ions on adjoining step
as this algorithm generates trajectories in both NVT ensembles edges.
(used for the surface simulations) and NPT ensembles (used
for the potential derivation), thus keeping the simulations con- RESULTS
sistent. The Nosé-Hoover parameters were set at 0.5 ps for both All three minerals have a rhombohedral crystal structure,
the thermostat and barostat relaxation times. In DL_POLY the which can be described by either a rhombohedral or hexagonal
shells are assigned a small mass and the choice of shell mass is axial system (Reeder 1990). As the two axial systems can make
a balance between keeping the mass as close to zero as pos- the description of the surfaces by their indices ambiguous, in
sible to ensure that no vibrational energy is exchanged with this work the hexagonal axial system and the resulting four
other species, notably hydrogen, and making the shells heavier, Miller-Bravais indices are used to describe each of the mineral
which allows a longer timestep during the simulations and hence surfaces. The pure end-members calcite CaCO3 and magnesite

computationally less expensive calculations. After a series of MgCO3 have the R3c space group, but the dolomite structure
comprises alternating planes of Mg and Ca ions (separated by
carbonate planes) in the c direction (Fig. 1), and as a result

TABLE 1. Potential model dolomite has the R3 space group, similar to calcite and magne-
Ion Core charge (e) Shell charge (e) Core-shell site, but with the loss of the c-glide. The experimental and cal-
interaction culated properties of the three minerals are listed in Table 2,
(eVÅ-2)
Ca, Mg +2.000 from which one can see that not only the calculated structural
C +1.135 but also the elastic properties agree well with the experimen-
H +0.400
carbonate oxygen (O) +0.587 –1.632 507.400000
tally determined values, even though the structure of dolomite
water oxygen (Ow) +1.250 –2.050 209.449602 and the elastic moduli for magnesite and dolomite were not
used in the derivation of the potential parameters. The poten-
Ion pair Buckingham potential
A (eV) r (Å) C (eVÅ6) tial model for MgCO3 was not only derived from the structure
Ca-O 1550.0 0.29700 0.0 of the magnesite crystal, but the parameters of the MgCO3 crys-
Ca-Ow 1186.6 0.29700 0.0 tal and its interactions with water were also fitted to the ex-
Mg-O 1092.2 0.27926 0.0
Mg-Ow 2290.0 0.22000 0.0
H-O 396.3 0.23000 0.0
H-Ow 396.3 0.25000 10.0
O-O 16372.0 0.21300 3.47
O-Ow 12533.6 0.21300 12.09

Lennard-Jones potential
A (eVÅ12) B (eVÅ6)
Ow-Ow 39344.98 42.15

Morse potential
D (eV) a (Å–1) r0 (Å)
C-O 4.710000 3.80000 1.18000
H-Ow 6.203713 2.22003 0.92376

Three-body potential
k (eV/rad2) Q0
Ocore-C-Ocore 1.69000 120.000000
H-Owshell-H 4.19978 108.693195

Four-body potential
k (eV/rad2) Q0
C-Ocore-Ocore-Ocore 0.11290 180.0

Coulombic subtraction (%) FIGURE 1. Bulk structure of dolomite, showing alternating planes
H0.4+-O0.8- 50 of calcium, carbonate, and magnesium (Ca = gray, Mg = black, CO3 =
H0.4+-H0.4+ 50 white three-spoked wheel).
682 DE LEEUW: MODELING DOLOMITE AND MAGNESIAN CALCITES


perimental enthalpies for the processes described in Equations {1120} surfaces of calcite, dolomite, and magnesite to investi-
1 and 2 below, which were calculated at +304.9 kJ/mol and gate how the relative stabilities of the surfaces depend on in-
+23.3 kJ/mol respectively [cf. exp. +311.5 and +35.2 kJ/mol creasing numbers of Mg ions in the structures and also how
(Lide 2000)]. The energies of the hydrated ions (Mg2+, Ca2+, they alter as water is adsorbed at the surface. To this end the
and CO32–) were obtained from Molecular Dynamics simula- surface energies of both dry and wet surfaces were calculated,
tions at constant pressure and temperature, where each ion was where the latter had a monolayer of water interacting with the
surrounded by 255 water molecules in a box, which was peri- surface [in this work a monolayer is defined to mean a surface
odically repeated in three dimensions. coverage of one water molecule per surface calcium or car-
bonate group, which was found to be the preferred configura-

CaCO3(s) + Mg(g)
2+
Æ MgCO3( s ) + Ca (2g+) (1) tion on the dominant {1014} surface (de Leeuw et al. 1999)].
The water molecules were placed above the surface, after which
CaCO3( s ) + Mg(2aq+ ) Æ MgCO3( s ) + Ca (2aq+ ) (2) both the surface and the adsorbing water molecules were al-
lowed to fully relax using a Newton Raphson energy minisation
The accuracy of the CaCO3-H2O interactions was verified technique (Watson et al. 1996). A host of initial configurations
by modeling the hydrous calcium carbonate phase ikaite, which for the water molecules was calculated, e.g., adsorbed to sur-
is a crystalline calcium carbonate hexahydrate, usually found face cations or O atoms or rotated in various positions, to en-
only at very low temperatures, and often in the presence of sure as far as possible that the lowest energy configuration had
calcite inhibititors (Bischoff et al. 1993; Clarkson et al. 1992). been obtained. In all cases the water molecules show a clear
Although thermodynamically stable with respect to the aque- preference for coordination by O ions, but where possible also
ous calcium and carbonate ions (DG = –37 kJ/mol), ikaite is hydrogen bonding to surface O atom ions. Where the same sur-
less stable than the anhydrous form calcite [by about 10 kJ/ faces were calculated for the three different minerals, the simu-
mol (Lide 2000)] and it is generally accepted as one of the pre- lations were performed on the same number of formula units
cursor phases in the precipitation of calcite from aqueous solu- in the simulation cells, so that the surface areas and surface
tions (Bischoff et al. 1993; Ito 1998). In an earlier work, good energies for the three minerals could be compared directly. The
agreement was found between the experimental and calculated simulation cells typically consisted of 40–80 unit cells, each
structural parameters of ikaite (de Leeuw and Parker 1997), containing six MCO3 formula units, giving a maximum of 3840
and a change in enthalpy for the dissocation of ikaite into cal- species per simulation cell. The surface energy g (Eq. 3) of a
cite and water of 47 kJ/mol per water molecule was calculated, particular surface is a measure of the stability of that surface
which compared very well with experimental values of 47–50 with a low positive value indicating high stability.
kJ/mol (Bischoff et al. 1993). In addition, when hydration of
Us - U b
the surfaces was taken into account, the experimental morpholo- g= (3)
gies of calcite and aragonite were reproduced accurately, which A
shows that the relative stabilities of the various surfaces were where Us is the energy of the surface block of the crystal, Ub is
calculated accurately (de Leeuw and Parker 1998a). Further- the energy of an equal number of atoms of the bulk crystal and
more, calcite surface features which had been observed by A is the surface area. When the surfaces are hydrated, the sur-
atomic force microscopy (AFM) were reproduced only when face energy is calculated as follows:
adsorbed water molecules were included in the surface calcu-
U h - ( U b + nU H 2 O )
lations (de Leeuw and Parker 1997), all of which indicate that gw = (4)
the water-solid interactions are modeled reliably. A
where Uh is the energy of the relaxed surface with n adsorbed
Structures and relative stabilities of planar surfaces water molecules and UH2O is the energy of a liquid water mol-
Static lattice energy minimization techniques were first ecule, which is the sum of the calculated self-energy of an iso-
– – –
employed to model the {1014}, {0001}, {1010}, {1011}, and lated water molecule and the condensation energy of water,
which were calculated using MD simulations at –43.0 kJ/mol
(de Leeuw and Parker 1998b). This is in excellent agreement
with the experimental value [–43.4 kJ/mol (Duan et al. 1995)],
TABLE 2. Calculated and experimental properties of calcite, mag- especially as the potential parameters were not fitted to the
nesite, and dolomite condensation energy.
Property Calcite Magnesite Dolomite The calculated surface and hydration energies for the vari-
exp. calc. exp. calc. exp. calc. ous surfaces are given in Tables 3 and 4 respectively. The
V (Å3) 367.8 348.0 279.1 260.1 323.6 300.9 –
(–5.4%) (–6.8%) (–7.0%)
{0001} and {1011} surfaces are dipolar and have more than
a,b (Å) 4.990 4.796 4.633 4.444 4.810 4.640 one surface termination, being either terminated by cations or
c (Å) 17.061 17.470 15.016 15.210 16.010 16.139 by carbonate groups. For the calcite and magnesite crystals, I
c/a 3.419 3.643 3.241 3.422 3.328 3.478
Bulk 73.5* 80 117† 123 91‡ 97 need only consider two surfaces, but in the case of dolomite
modulus there are two cation-terminated planes, either Ca or Mg, and in
(GPa) addition two carbonate terminated planes, one with Ca and the
* Redfern and Angel (1999).
† Ross (1997).
other with Mg ions in the second surface layer. All of these
‡ Martinez et al. (1996). different surface terminations need to be considered to calcu-
DE LEEUW: MODELING DOLOMITE AND MAGNESIAN CALCITES 683

late the overall stability of these dipolar surfaces and to deter- interactions to surface O ions at approximately 1.8 Å. Molecu-
mine which surface termination will be the dominant plane. lar dynamics simulations of a monolayer of water on the cal-

The dipoles are removed by the creation of surface vacancies, cite {1014} surface at 300 K showed that there is no surface
where the resulting planes are 50% vacant with respect to ei- diffusion of the surface water molecules, which remain adsorbed
ther surface cations or carbonate groups (Tasker 1979), which in the same pattern as found here (de Leeuw et al. 1999). On
is why the nature of the underlying second surface layer can- the magnesite surface, the water molecules are adsorbed in an
not be ignored, as its ions are exposed underneath the carbon- equally regular pattern (Mg-Ow distance of 1.85 Å) but rotated
ate vacancy sites. with respect to the calcite surface. Due to the smaller area of
As can be seen from Table 3, the surface energies of the the surface unit cell, the rotated adsorption pattern ensures that

{1014} surfaces of the three minerals before hydration increase
with the increasing number of surface Mg ions, from g = 0.59
J/m2 for calcite to g = 0.64 J/m2 for dolomite and g = 0.76 J/m2
for magnesite, showing that the non-hydrated purely Ca sur-
face of calcite is the most stable and the purely Mg surface of
magnesite the least stable. However, once hydrated, the trend
is reversed with magnesite becoming the most stable surface (g
= 0.02 J/m2), followed by dolomite (g = 0.24 J/m2), and finally
calcite as the least stable surface (g = 0.33 J/m2). This indicates
that under aqueous conditions, formation of a structure con-
taining a complete Mg surface is energetically more favorable
than formation of the mixed Mg-Ca surface in dolomite. The
difference in stabilization of the three surfaces is due to the
dissimilar patterns of adsorbed water molecules (see Fig. 2).
On the calcite surface, the water molecules form a regular her-
ringbone pattern, closely coordinated to the surface. The water
molecules are coordinated by their O ions to surface Ca ions at
a distance of 2.4 Å, in good agreement with earlier calcula-
tions (de Leeuw and Parker 1997) and subsequent X-ray
reflectivity measurements by Fenter et al. (2000), who mea-
sured a Ca-OHx distance of 2.50 ± 0.12 Å (where x = 1, 2, or
3). In addition, the water molecules form hydrogen-bonded

TABLE 3. Surface energies of calcite, magnesite and dolomite (Jm–2)


Surface Calcite Magnesite Dolomite
dry wet dry wet dry wet

{101 4} 0.59 0.33 0.76 0.02 0.64 0.24
{0001}Ca 0.97 0.76 – – 1.04 0.83
{0001}Mg – – 1.41 0.82 1.29 0.77
{0001}CO3 (Ca) 0.99 0.70 – – 1.19 0.88
{0001}CO3 (Mg) – – 1.41 0.59 1.21 0.47
{1010}
– 0.97 0.82 1.26 0.80 1.13 0.84
{101–1}Ca 1.23 0.88 – – 1.28 0.83
{101–1}Mg – – 1.61 0.74 1.17 0.69
{101–1}CO3 (Ca) 1.24 0.95 – – 1.33 1.10
{101 1}CO3 (Mg) – – 1.42 0.99 1.30 0.79
{112–0} 1.39 0.80 1.90 0.70 1.28 0.82

TABLE 4. Hydration energies of calcite, magnesite, and dolomite


surfaces (kJmol–1)
Surface Calcite Magnesite Dolomite

{101 4} –73.7 –114.5 –85.8
{0001}Ca –68.6 – –67.6
{0001}Mg – –104.7 –101.9
{0001}CO3 (Ca) –67.2 – –78.5
{0001}CO3 (Mg) – –127.9 –127.5
{1010}
– –81.9 –137.4 –107.9
{101–1}Ca –54.2 – –113.7 –
FIGURE 2. Plan view of the hydrated {1014} surfaces of (a) calcite,
{101–1}Mg – –164.8 –118.2
{101–1}CO3 (Ca) –74.9 – –80.1
(b) magnesite, and (c) dolomite, showing the different adsorption
{101 patterns water molecules (MCO3 as framework: Ca = gray, Mg = black,
– 1}CO3 (Mg) – –102.8 –122.5
{112 0} –130.5 –184.5 –103.1 CO3 = white, water space filling: Owater = black, H = white).
684 DE LEEUW: MODELING DOLOMITE AND MAGNESIAN CALCITES


the water molecules not only form hydrogen bonds to surface of the calcite {1014} surface, despite the presence of Mg ions
O ions, but they are now also capable of forming a network of in the dolomite surface layer. On the other two mixed sur-
– –
hydrogen bonds inter-molecularly between the water molecules, faces, {1010} and {1120}, the hydration energies are dissimi-

which is not possible on calcite because of the larger lattice lar to both calcite and magnesite, and hydration of the {1120}
spacing. However, on the mixed Mg-Ca dolomite surface, the surface releases only 103 kJ/mol, compared to 131 and 185 kJ/
smaller Mg ions have receded into the surface with the Ca ions mol for calcite and magnesite, respectively. The hydrated

protruding above the surface. As a result the Mg ions are less {1120} surfaces of magnesite, calcite, and dolomite are shown
accessible to adsorbing water molecules and the regular pat- in Figure 3. The magnesite surface is virtually bulk-terminated,
tern of water molecules found in calcite and magnesite is dis- without detectable relaxation of the carbonate groups, and the
turbed. The water molecules adsorb in pairs, where there is water molecules are closely coordinated to the surface Mg ions
hydrogen-bonded coordination between the two paired water in addition to interacting by their hydrogen atoms to O ions of
molecules, but no extensive intermolecular network as was neighboring water molecules and in the surface. In the calcite
found in magnesite. Furthermore, because of the lesser acces- surface, however, although the cation sublattice remains unaf-
sibility of the Mg ions, the water is less strongly coordinated to fected, extensive rotation of the carbonate groups occurs in the
surface cations and as a result the hydrated dolomite surface is first three surface layers, and half the topmost carbonate groups
less stable than its pure calcium or magnesium counterpart. now lie almost flat in the plane. Calcite has a greater lattice

Despite the strong affinity of Mg for water, the surface energy spacing (surface area {1120} = 48.4 Å2) than magnesite (sur-
2
of the hydrated dolomite plane is similar to calcite rather than face area = 39.1 Å ), and even after rotation of the surface car-
magnesite. bonate groups, the sub-surface Ca ions are still accessible to
In general, the other dehydrated surfaces follow the same the water molecules, which bond to the Ca ions by their O at-

trend as the {1014} surfaces, where the surface energies for oms. These water molecules also form some hydrogen-bonded
magnesite are higher than those for calcite, but now they re- interactions to surface O ions, although to a lesser extent than
main similar to calcite even when the surfaces are hydrated, on magnesite, and hence the hydration energy is lower for cal-

which shows that the extremely low surface energy of the hy- cite. The carbonate groups in the dolomite {1120} surface (sur-

drated magnesite {1014} surface is a special case, due to the face area = 43.3 Å2) rotate to a limited extent in the surface
fact that the lattice spacing enables the formation of an unusu- layer. However, this partial rotation makes the Mg ions far less
ally stable network of hydrogen-bonded coordinations between accessible to the water molecules than in the magnesite sur-
the water molecules and to the surface (described above). Of face. In addition, the smaller lattice spacing results in less ad-
– – –
the dolomite surfaces, the {1014}, {1010}, and {1120} planes sorption of water to the Ca ions, and both effects combine to

have mixed Mg/Ca layers, while the {0001} and {1011} sur- lessen the hydration energy of this dolomite surface with re-
faces contain alternating planes of Mg or Ca ions, separated by spect to the pure end-members.
carbonate layers. When I compare the surface energies of the The hydration energies of the dolomite {0001} planes are
dolomite surfaces, I see that for the dehydrated {0001} surface very similar to their counterparts in magnesite and calcite, but

the planes where the first cation layer consists of Mg are less the hydration pattern on the {1011} planes is again disturbed.
stable than the planes terminated by Ca ions in the first cation The Ca-terminated dolomite plane has a much higher hydra-

layer, while the reverse is the case for the {1011} surface, where tion energy than calcite while hydration of the Mg-terminated
the Mg-terminated surfaces are more stable. Once hydrated, magnesite plane releases far more energy than the dolomite
the Mg-terminated planes are the more stable planes for both plane, due to the loss of an extensive intermolecular network
– –
{0001} and {1011} surfaces. Similar to the {1014} surfaces, of hydrogen-bonded interactions between the water molecules

the surface energy of the dehydrated dolomite {1010} surface on the smaller magnesite surface (surface area = 46.5 Å2 com-
(1.13 J/m2) falls between that for calcite (0.97 J/m2), and that pared to 51.5 Å2 on dolomite). Rotation of the carbonate groups

for magnesite (1.26 J/m2), but the dehydrated dolomite {1120} on the carbonate-terminated surfaces leads to a similar hydra-
surface is more stable than either its calcite or magnesite coun- tion energy for dolomite compared to calcite, but not compared

terpart. Once hydrated, some of the dolomite surfaces are less to its magnesite counterpart. Contrary to the {1120} surface,
stable than the equivalent magnesite surface, but the Mg-ter- where rotation of the surface carbonate groups made the sur-

minated {0001} and {1011} planes have lower surface ener- face Mg ions less accessible to the adsorbing water molecules,

gies than their magnesite counterparts. Again, I see that the the surface carbonate groups on the carbonate-terminated {101
mixed Ca/Mg surfaces are less stabilized by hydration than the 1} planes rotate away from the underlying cations. These cat-
purely Mg surfaces, due to the strong interactions between sur- ions are now more accessible to the water molecules and hence,
face Mg ions and the water molecules and the reduced accessi- this time, the hydration energy for dolomite is larger than for

bility of the Mg ions to the water molecules, once Ca is magnesite, where again bulk termination of the {1011} sur-
incorporated in the surfaces as well. face occurs without rotation of surface carbonate groups.
Comparing the hydration energies for the different miner- In addition to the fully ordered dolomite structure, a collec-
als (Table 4), the hydration energies for the magnesite surfaces tion of less-ordered magnesian calcites was studied as well,
are generally larger than those for calcite. The weaker binding where Mg ions are introduced at various depths in the calcite

of water to the dolomite {1014} surface, compared to the mag- surfaces, leading to substitutions of Ca by Mg ions of between

nesite {1014} surfaces, is shown quantitatively in the hydra- 6 to 25%. The surface and hydration energies of some surfaces
tion energy. For dolomite this resembles the hydration energy of these magnesian calcites are collected in Table 5, where only
DE LEEUW: MODELING DOLOMITE AND MAGNESIAN CALCITES 685

the calculated values are shown for the planes that are directly
comparable to the dolomite surfaces. Magnesium is either in-
troduced in the topmost cation layer or in the second layer in
the same ratio as in dolomite. In a previous work (de Leeuw
and Parker 2000) we find that Mg introduced to the surface in
the presence of water, does not segregate into the bulk mineral,
but only dissolves into the bulk when the surface water is re-
moved. This study is primarily concerned with the possible
dolomitization of existing calcite and in the absence of diffu-
sion of Mg into the bulk, I found that introducing Mg deeper
into the structure (third and fourth sub-surface layers) did not
significantly affect the structures and energies compared to pure
calcite. Hence, the structures of the disordered magnesian cal-
cites, given in Table 5, all consist of an underlying bulk calcite
structure, but with a dolomite-type ordering of Mg and Ca ions
in the topmost surface layer(s). These magnesian calcites may
resemble, on a much smaller scale, some dolomite-like miner-
als, which are found to be richer in Ca than pure dolomite and
have a superstructure of larger lattice dimensions (Schubel et
al. 2000).
When comparing the surface energy of the dehydrated

{1014} surface of the magnesian calcite with those for calcite,
magnesite, and dolomite, this disordered structure is equally
stable as the fully ordered dolomite structure despite the un-
derlying calcite lattice. Hydration of the magnesian calcite leads
to a very similar pattern of adsorption of water molecules as on
the magnesite surface (Fig. 2b), even though the distribution
of Mg ions in the surface layer of the disordered material is the
same as in dolomite. However, because the bulk lattice is that
of calcite rather than dolomite, the surface Mg ions are not
obscured by the Ca ions to a similar extent as in dolomite, and
they thus remain accessible to the water molecules. However,
due to the larger lattice spacing of the magnesian calcite com-
pared to magnesite, the extensive network of hydrogen-bonded
interactions between the adsorbing water molecules is not
present in the magnesian calcite, which is reflected in the
smaller hydration energy for the disordered material compared
to that for magnesite (Table 5).
All of the other surface energies for the magnesian calcites
are lower than those for dolomite, both when dehydrated and
when water is present at the surface. As the only difference
between the surfaces of dolomite and magnesian calcite is the
increased lattice spacing of the latter (due to the underlying

TABLE 5. Surface and hydration energies of dolomite-like surfaces


of magnesian calcites
Surface Surface energies (Jm–2) Hydration energies (kJmol–1)
Dry Wet

{101 4} 0.65 0.25 –90.5
{0001}Ca 0.73 0.54 –65.6
{0001}Mg 1.03 0.59 –96.8
– {0001}CO3 (Ca) 0.87 0.42 –79.4
FIGURE 3. Elevation view of the {1120} surfaces of (a) magnesite, {0001}CO3 (Mg) 0.96 0.42 –83.2
(b) calcite, and (c) dolomite, showing little relaxation of the surface {1010}
– 0.98 0.72 –108.8
cations, but rotation of the carbonate groups in calcite and dolomite, {101–1}Ca 1.11 0.52 –111.9
{101–1}Mg 1.01 0.73 –75.8
together with the different patterns of adsorption of the water molecules {101–1}CO3 (Ca) 1.02 0.75 –90.4
(MCO3 as framework: Ca = gray, Mg = black, CO3 = white, water {101
– 1}CO3 (Mg) 1.08 0.65 –117.9
space filling: Owater = black, H = white). {112 0} 1.25 0.78 –111.4
686 DE LEEUW: MODELING DOLOMITE AND MAGNESIAN CALCITES

bulk calcite structure), rather than the cation distribution in the


surface, the stability of the surfaces is clearly influenced to a
large extent by the more open structure of the magnesian cal-
cite. This open structure has a twofold effect; first it results in FIGURE 4. Schematic representation of the modeled growth process
a more effective relaxation of the Mg-containing surface lay- at the calcite step edges, where the black shaded box is the newly
ers, hence lowering the surface energies of the dehydrated sur- incorporated CaCO3 or MgCO3 unit.
faces and, secondly, it enhances the accessibility of the surface

Mg ions to the adsorbing water molecules, hence further low- {1014} surface of dolomite is such that if this surface is to
ering the surface energies of the hydrated surfaces. As the hy- grow from the calcite steps, the edge of each growth step should
dration energies of the magnesian calcite surfaces are actually consist of a regular array of alternating CaCO3 and MgCO3
either similar or less than those for the equivalent surfaces in groups.
dolomite (Table 5), the first effect is the more important in de- I first calculated the energetics of adding isolated CaCO3
termining the stabilities of the surfaces. and MgCO3 units at either step (step 1 in Fig. 4), introducing
It is clear from the above calculations of surface energies of the first kink sites at the perfect step edges, then I calculated
dolomite, calcite, and magnesite that many of the dolomite sur- the addition of CaCO3 and MgCO3 units next to other CaCO3
faces in an aqueous environment are thermodynamically less units on a growing edge (steps 2 and 3 in Fig. 4), and finally I
stable than the surfaces of its constituent carbonates calcite and calculated the addition of the last unit into a step, annihilating
magnesite. In addition, a disordered magnesian calcite is also all kink sites and recreating a perfect step edge (step 4 in Fig.
energetically preferred over the ordered dolomite structure. As the 4). Figure 6 shows graphs of the energetics of these steps for
magnesian calcites were modeled on a calcite bulk structure, the
next step was to investigate the growth of such materials, by mod-

eling the incorporation of MgCO3 into the calcite {1014} surface.

MgCO3 growth at calcite steps


Although diffusion and Ca-Mg cation exchange may play a
role in dolomite formation, an alternative mechanism is by the
incorporation of Mg in a growing calcite crystal. Calcite crys-
tal growth occurs at steps. Hence, in contrast to the above com-
parison of the planar surfaces of calcite, magnesite, dolomite,
and magnesian calcite, the next task was to investigate the up-

take of Mg2+ ions at steps on the dominant {1014} surface.
Molecular dynamic simulations were used, enabling inclusion
of temperature in the calculations, to calculate the energetics
of incorporation of Mg ions in a growing calcite crystal (at 300
K). Each step in the growth process, namely the incorporation
of a new MgCO3 or CaCO3 unit at the growing step, was mod-
eled by a full molecular dynamic simulation including relax-
ation of the solid surface and the water molecules. Figure 4
shows a schematic representation of the modeled growth pro-
cess, where the dark shaded box is the newly incorporated
CaCO3 or MgCO3 unit.
Calcite growth is experimentally found to occur from two

monatomic steps on the {1014} surface (e.g., Gratz et al. 1993;
Liang et al. 1996a, 1996b). The absorption of Mg ions was
therefore studied at these two growth steps, to investigate pos-
sible dolomitization of unconsolidated calcium carbonates, in-
cluding solvent effects to mimic sedimentary conditions. One
of the growth steps is acute, i.e., the carbonate groups on the
edge of the step overhang the plane below the step with a cal-
culated angle between relaxed step wall and plane of 80∞ [cf.
exp. 78∞ (Park et al. 1996)] (Fig. 5a), while the other step is
obtuse, i.e., the carbonate groups on the step edge lean back
with respect to the plane below, forming an angle between re-
laxed step wall and surface plane of 105∞ [cf. exp. 102∞ (Park
et al. 1996)] (Fig. 5b). For a mixed Mg-Ca surface to be formed, FIGURE 5. Side view of the (a) acute and (b) obtuse calcite growth
it should be energetically feasible to incorporate Mg ions at Ca –
steps, showing the {1014} planes, offset from each other by monatomic
carbonate steps and calcium ions at steps decorated with mag- steps (MCO3 as framework: Ca = gray, Mg = black, CO3 = white, water
nesium carbonate. The arrangement of Ca and Mg ions in the space filling: Owater = black, H = white).
DE LEEUW: MODELING DOLOMITE AND MAGNESIAN CALCITES 687

the addition of either a MgCO3 or a CaCO3. In all cases (steps by a second MgCO3 unit, and finally completing the edge with
1–4 in Fig. 4), addition of an MgCO3 unit is thermodynami- another CaCO3 unit. The enthalpies of these processes are also
cally favorable, whether incorporated onto a complete edge displayed in Figure 8, which shows that there is a large penalty
(step 1) or next to a CaCO3 corner site (steps 2–4). However, in energy in step 2 of the process. On the obtuse step, incorpo-
the addition of CaCO3 units is energetically unfavorable when ration of this CaCO3 group costs 34 kJ/mol, while adding an-
incorporated onto a complete edge or next to an isolated CaCO3 other MgCO3 unit releases 51 kJ/mol, a difference of 85 kJ/
unit. It is only when two (acute step) or three (obtuse step) mol. The same processes on the acute step are energetically
neighboring CaCO3 units are already present that the addition even more divergent, requiring and releasing 19 and 82 kJ/mol,
of a further CaCO3 unit is an exothermic process. In the case of
the obtuse edge this means that only the annihilation of the
kink sites and re-formation of a complete edge is energetically
favorable. As these calculations suggest that the growth of
MgCO3 onto the calcite steps is energetically preferred over
CaCO3 growth, the incorporation of further MgCO3 units next
to the first and subsequent MgCO3 units was also calculated, in
effect growing a complete MgCO3 edge onto both calcite steps.

Figure 7 shows a plan view of the {1014} surface with com-
TOP
pletely Mg-decorated acute step edges, which are offset from
each other in height by one monatomic layer (as found experi-
mentally by Hillner et al. 1993), where for clarity the water
molecules are not shown. There is some relaxation of the ions
MIDDLE
on the step edge, notably the carbonate groups, which have
rotated to best accommodate the Mg ions, but the Ca ions and
carbonate groups on the terraces remain virtually in their bulk-
BOTTOM
terminated positions. It is clear from the graph in Figure 8 that
further growth of MgCO3 units onto both edges is also ener-
getically very favorable, and far more exothermic than CaCO3
growth (energies shown in Fig. 6). However, the formation of
dolomite requires the presence of both Mg and Ca at the sur-

face and the arrangement of the Mg and Ca ions in the {1014}
surface of dolomite is such that I need to incorporate alternat-
ing CaCO3 and MgCO3 units at both growth steps to create a
surface that has the dolomitic cation arrangement. The calcu- –
lated energies for this alternating arrangement of Mg and Ca FIGURE 7. Plan view of three calcite {1014} planes (top, middle,
and bottom), offset from each other by acute steps, showing fully
ions at the growing edges are displayed in Figure 8. As the
magnesium-decorated step edges, where for clarity the water molecules
incorporation of an MgCO3 unit at the perfect edge is the ener- are not shown (Ca = gray, Mg = black, CO3 = white).
getically more favorable process, this MgCO3 unit was taken
as a starting point, subsequently adding a CaCO3 unit, followed

FIGURE 6. Energies of incorporation of MgCO3 and CaCO3 units FIGURE 8. Energies of sequential growth of MgCO3 units at acute
onto the acute and obtuse calcite growth steps at the different stages and obtuse calcite steps, together with energies of growth of dolomite

of the growth process shown in Figure 4. edges at the acute and obtuse steps on the calcite {1014} surface.
688 DE LEEUW: MODELING DOLOMITE AND MAGNESIAN CALCITES

respectively, a difference in energy of over 100 kJ/mol. The faces is gcal:gmag:gdol:gdisordered = 17:1:12:12, clearly showing the
last two steps in the formation of a dolomite edge are exother- unusual stabilities of magnesite (which is also due to the ex-
mic, but in view of the large energies required to incorporate tensive network of hydrogen-bonded interactions between the
the first CaCO3 unit, especially in competition with Mg ions adsorbed water molecules) especially compared to calcite.
present in the solution, dolomite growth is not expected to oc- The hydration energies of the pure minerals are largest for
cur at ambient temperatures. For completeness sake, energies magnesite (at an average of 134 kJ/mol over all surfaces), fol-
of mixed Mg and Ca edges with less regular arrangements of lowed by dolomite and calcite (average 101 and 79 kJ/mol re-
the two cations were also calculated and it was found that once spectively). However, whereas the average hydration energy
MgCO3 and CaCO3 units had been incorporated at the obtuse of the pure dolomite falls between magnesite and calcite, the
step, there is a small energetic advantage (10 kJ/mol) for the average hydration energy for the disordered magnesian calcite
next MgCO3 unit to grow next to the CaCO3 group rather than is less at 94 kJ/mol. This lower average hydration energy for
next to the first MgCO3 unit. Hence, if a CaCO3 group did grow the disordered structure is partly due to the greater lattice spac-
next to the first MgCO3 unit, the next step would indeed lead to ing of magnesian calcite compared to both magnesite and do-
a dolomite-type arrangement on the obtuse edge. However, on lomite, which makes the formation of a network of
the acute edge, the reverse happens and the second MgCO3 hydrogen-bonded interactions between the adsorbed water
unit would preferentially grow next to the first MgCO3, rather molecules impossible, and partly due to the more effective re-
than next to the CaCO3 unit (by about 14 kJ/mol). In addition, laxation of the anhydrous disordered surfaces, which therefore
the last unit to be incorporated on both edges should be a CaCO3 gain less on hydration than the magnesite and dolomite sur-
group, but it is energetically far preferable to incorporate an- faces.
other MgCO3 unit (–150 and –205 kJ/mol for the acute and Molecular dynamics simulations of the competitive growth
obtuse edges, respectively) than another CaCO3 unit (–51 and of CaCO3 and MgCO3 at the two experimentally observed cal-
–22 kJ/mol for the acute and obtuse edges, respectively). Thus, cite steps show that, at all stages, incorporation of MgCO3 is
these molecular dynamics calculations of dolomite growth at more favorable than CaCO3, indicating that if enough Mg is
two experimentally observed calcite growth steps show that present a complete MgCO3 plane will form. When investigat-
growth of MgCO3 at the steps is preferred over incorporation ing the formation of dolomite-type layers from the steps, I find
of CaCO3 at all stages of the growth process and it would there- a small energetic advantage for an ordered arrangement of Ca
fore be very unlikely for an ordered structure like dolomite to and Mg ions on the obtuse step (10 kJ/mol), but not on the
be formed in this way. It is, however, clear that on thermody- acute step. As the obtuse step is experimentally found to grow
namic grounds MgCO3 growth on calcite will proceed very and dissolve more rapidly than the acute step (MacInnis and
easily, which explains why highly magnesian calcites occur Brantley 1992; Liang et al. 1996a, 1996b), this may show a route
widely in natural sediments. to at least partial dolomite formation. Even so, however, the incor-
poration of MgCO3 rather than CaCO3 is far more exothermic and
DISCUSSION hence, a mixed Mg-Ca carbonate would only form if Mg is present
The calculated structural and elastic properties of the three at very small concentrations.
minerals are in good agreement with experimental values, even
though bulk moduli were not used in the derivation of the po- ACKNOWLEDGMENTS
tential parameters. The relative enthalpies of formation of the N.H.dL. thanks NERC, grant no. NER/M/S/2001/00068, and the Royal
calcite and magnesite crystals were fitted to experimental en- Society, grant no. 22292, for financial support.
thalpies and are therefore reproduced well, which was a prerequi-
site for the calculations of the varying stabilities of the pure and
REFERENCES CITED
Baker, P.A. and Kastner, M. (1981) Constraints on the formation of sedimentary
disordered minerals and the energies of the growth processes. dolomite. Science, 213, 114–116.
The calculated surface energies for the three carbonate min- Bertram, M.A., Mackenzie, F.T., Bishop, F.C., and Bischoff, W.D. (1991) Influence
erals show that, in general the dolomite surfaces are thermody- of temperature on the stability of magnesian calcite. American Mineralogist,
76, 1889–1896.
namically more stable than their counterparts in magnesite, in Bischoff, J.L., Fitzpatrick, J.A., and Rosenbauer, R.J. (1993) The solubility and
an aqueous environment the dolomite surfaces are less stable stabilization of ikaite (CaCO3·6H2O) from 0 ∞C to 25 ∞C—environmental and
than the surfaces of its constituent carbonates calcite and mag- paleoclimatic implications for thinolite tufa. Journal of Geology, 101, 21–33.
Born, M. and Huang, K. (1954) Dynamical Theory of Crystal Lattices. Oxford Uni-
nesite. In addition, a disordered magnesian calcite is also cal- versity Press, Oxford.
culated to be energetically preferred over the ordered dolomite Brady, P.V., Krumhansi, J.L., and Papenguth, H.W. (1996) Surface complexation
clues to dolomite growth. Geochimica et Cosmochimica Acta, 60, 727–731.
structure compared to magnesite. The lesser stability of the hy- Clarkson, J.R., Price, T.J., and Adams, C.J. (1992) Role of metastable phases in the

drated {1014} dolomite surface is in part due to the strong Mg- spontaneous precipitation of calcium-carbonate. Journal of the Chemical Soci-
H2O interactions, which upon hydration lower the surface ety Faraday Transactions, 88, 243–249.
de Leeuw, N.H. and Parker, S.C. (1997) Atomistic simulation of the effect of mo-
energies of the magnesite surface far more than the calcite or lecular adsorption of water on the surface structure and energies of calcite sur-
dolomite planes, together with the fact that the Mg ions present faces. Journal of the Chemical Society Faraday Transactions, 93, 467–475.
in the dolomite surface are less accessible as a result of surface ———(1998a) Surface structure and morphology of calcium carbonate polymor-
phs calcite, aragonite and vaterite: an atomistic approach. Journal of Physical
relaxations and/or rotations of the carbonate groups. They are, Chemistry B, 102, 2914–2922.
however, more accessible in the magnesian calcites, due to the ———(1998b) Molecular-dynamics simulation of MgO surfaces in liquid water
using a shell-model potential for water. Physical Review B, 58, 13901–13908.
larger lattice spacing of the underlying calcite structure. The ———(2000) Modelling absorption and segregation of magnesium and cadmium

ratio of surface energies of the dominant hydrated {1014} sur- ions to calcite surfaces: introducing MgCO3 and CdCO3 potential models. Jour-
DE LEEUW: MODELING DOLOMITE AND MAGNESIAN CALCITES 689

nal of Chemical Physics, 112, 4326–4333. nite and dolomite at high pressure and high temperature: Evidence for dolomite
de Leeuw, N.H., Parker, S.C., and Harding, J.H. (1999) Molecular Dynamics simu- breakdown to aragonite and magnesite. American Mineralogist, 81, 611–624.
lation of crystal dissolution from calcite steps. Physical Review B, 60, 13792– Nordeng, S.H. and Sibley, D.F. (1994) Dolomite stoichiometry and Ostwald step
13799. rule. Geochimica et Cosmochimica Acta, 58, 191–196.
Deer, W.A., Howie, R.A., and Zussman, J. (1992) An Introduction to the Rock- Nosé, S. (1984) A unified formulation of the constant temperature molecular-dy-
Forming Minerals, 641 p. Longman Group, Harlow, U.K. namics methods. Journal of Chemical Physics, 81, 511–519.
Dick, B.G. and Overhauser, A.W. (1958) Theory of the dielectric constants of alkali Oliver, P.M., Parker, S.C., Egdell, R.G., and Jones, F.H. (1996) Computer simula-
halide crystals. Physical Review, 112, 90–103. tion of the surface structures of WO3. Journal of the Chemical Society Faraday
Duan, Z., Moller, N., and Weare, J.H. (1995) Equation of state for the NaCl-H2O- Transactions, 92, 2049–2056.
CO2 system—prediction of phase-equilibria and volumetric properties. Paquette, J., Vali, H., and Mountjoy, E. (1999) Novel TEM approaches to imaging
Geochimica et Cosmochimica Acta, 59, 3273–3283. of microstructures in carbonates: Clues to growth mechanisms in calcite and
Fenter, P., Geissbuehler, P., DiMasi, E., Srajer, G., Sorensen, L.B., and Sturchio, dolomite. American Mineralogist, 84, 1939–1949.
N.C. (2000) Surface speciation of calcite observed in situ by high-resolution X- Park, N.-S., Kim, M.-W., Langford, S.C., and Dickinson, J.T. (1996) Atomic layer
ray reflectivity. Geochimica et Cosmochimica Acta, 64, 1221–1228. wear of single-crystal calcite in aqueous solution scanning force microscopy.
Ferry, J.M. (1996) Three novel isograds in metamorphosed siliceous dolomites from Journal of Applied Physics, 80, 2680–2686.
the Ballachulish aureole, Scotland. American Mineralogist, 81, 485–494. Parker, S.C. and Price, G.D. (1989) Computer modelling of phase transitions in
———(2000) Patterns of mineral occurrence in metamorphic rocks, American minerals. Advances in Solid State Chemistry, 1, 295–327.
Mineralogist, 85, 1573–1588. Pavese, A., Catti, M., Parker, S.C., and Wall, A. (1996) Modelling of the thermal
———(2001) Calcite inclusions in forsterite. American Mineralogist, 86, 773–779. dependence of structural and elastic properties of calcite, CaCO3. Physics and
Fiquet, G., Guyot, F., and Itie, J.P. (1994) High-pressure X-ray-diffraction study of Chemistry of Minerals, 23, 89–93.
carbonates–MgCO3, CaMg(CO3)2, and CaCO3. American Mineralogist, 79, 15– Pokrovsky, O.S., Schott, J., and Thomas, F. (1999a) Processes at the magnesium-
23. bearing carbonates/solution interface. I. A surface speciation model for magne-
Fisler, D.K. and Cygan, R.T. (1999) Diffusion of Ca and Mg in calcite. American site. Geochimica et Cosmochimica Acta, 63, 863–880.
Mineralogist, 84, 1392–1399. ———(1999b) Dolomite surface speciation and reactivity in aquatic systems.
Fisler, D.K., Gale, J.D., and Cygan, R.T. (2000) A shell model for the simulation of Geochimica et Cosmochimica Acta, 63, 3133–3143.
rhombohedral carbonate minerals and their point defects. American Mineralo- Reeder, R.J. (1990) Crystal chemistry of the rhombohedral carbonates. In Reeder,
gist, 85, 217–224. R.J., Ed., Carbonates: Mineralogy and Chemistry, 11, 1–48. Reviews in Miner-
Forester, T.R. and Smith, W. (1995) W. DL_POLY user manual. CCLRC, Daresbury alogy, Mineralogical Society of America, Washington, D.C.
Laboratory, Daresbury, Warrington, U.K. Redfern, S.A.T. and Angel, R.J. (1999) High-pressure behaviour and equation of
Gale, J.D. (1997) GULP: a computer program for the symmetry-adapted simulation state of calcite, CaCO3. Contributions to Mineralogy and Petrology, 134, 102–
of solids. Journal of the Chemical Society Faraday Transactions, 93, 629–637. 106.
Gratz, A.J., Hillner, P.E., and Hansma, P.K. (1993) Step dynamics and spiral growth Ross, N.L. (1997) The equation of state and high-pressure behavior of magnesite.
on calcite. Geochimica et Cosmochimica Acta, 57, 491–495. American Mineralogist, 82, 682–688.
Hillner, P.E., Manne, S., Hansma, P.K., and Gratz, A.J. (1993) Atomic Force Micro- Ross, N.L. and Reeder, R.J. (1992) High-pressure structural study of dolomite and
scope: A new tool for imaging crystal growth processes. Faraday Discussions, ankerite. American Mineralogist, 77, 412–421.
95, 191–197. Schubel, K.A., Elbert, D.C., and Veblen, D.R. (2000) Incommensurate c-domain
Hoover, W.G. (1985) Canonical dynamics: equilibrium phase-space distributions. superstructures in calcian dolomite from the Latemar buildup, Dolomites, North-
Physical Review A, 31, 1695–1697. ern Italy. American Mineralogist, 85, 858–862.
Ito, T. (1998) Factors controlling the transformation of natural ikaite from Shiowakka, Sibley, D.F., Nordeng, S.H., and Borkowski, M.L. (1994) Dolomitization kinetics
Japan. Geochemical Journal, 32, 267–273. in hydrothermal bombs and natural settings. Journal of Sedimentary Research
Katz, A. and Matthew, A. (1977) The dolomitization of CaCO3: An experimental Section A-Sedimentary Petrological Processes, 64, 630–637.
study at 252–295∞C. Geochimica et Cosmochimica Acta, 41, 297–308. Tasker, P.W. (1979) The surface energies, surface tensions and surface structure of
Kessels, L.A., Sibley, D.F., and Nordeng, S.H. (2000) Nanotopography of synthetic the alkali halides. Philosophical Magazine A, 39, 119–136.
and natural dolomite crystals. Sedimentology, 47, 173–186. Titiloye, J.O., de Leeuw, N.H., and Parker, S.C. (1998) Atomistic simulation of the
Krauskopf, K.B. and Bird, D.K. (1995) Introduction to Geochemistry. McGraw– differences between calcite and dolomite surfaces. Geochimica et Cosmochimica
Hill Inc. New York, 79. Acta, 62, 2637–2641.
Letargo, C.M.R., Lamb, W.M., and Park, J.S. (1995) Comparison of calcite plus Verlet, L. (1967) Computer ‘experiments’ on classical fluids. I. Thermodynamical
dolomite thermometry and carbonate plus silicate equilibria —constraints on properties of Lennard-Jones molecules. Physical Review, 159, 98–103.
the conditions of metamorphism of the Llano uplift, central Texas, USA. Ameri- Watson, G.W., Kelsey, E.T., de Leeuw, N.H., Harris, D.J., and Parker, S.C. (1996)
can Mineralogist, 80, 131–143. Atomistic simulation of dislocations, surfaces and interfaces in MgO. Journal
Liang, Y., Baer, D.R., McCoy, J.M., Amonette, J.E., and LaFemina, J.P. (1996a) of the Chemical Society Faraday Transactions, 92, 433–438.
Dissolution kinetics at the calcite-water interface. Geochimica et Cosmochimica Wenk, H.R. (1993) Partially disordered dolomite—microstructural characterization
Acta, 60, 4883–4887. of Abu-Dhabi Sabkha carbonates. American Mineralogist, 78, 769–774.
Liang, Y., Lea, A.S.,
– Baer, D.R., and Engelhard, M.H. (1996b) Structure of the cleaved Williams, Q., Collerson, B., and Knittle, E. (1992) Vibrational-spectra of magnes-
CaCO3 (1014) surface in an aqueous environment. Surface Science, 351, 172– ite (MgCO3) and calcite-III at high-pressures. American Mineralogist, 77, 1158–
182. 1165. –
Lide, D.R. (2000) CRC Handbook of Chemistry and Physics. CRC Press, Boca Wright, K., Cygan, R.T., and Slater, B. (2001) Structure of the (1014) surfaces of
Raton. calcite, dolomite and magnesite under wet and dry conditions. Physical Chem-
Liebermann, O. (1967) Synthesis of dolomite. Nature, 213, 241–245. istry and Chemical Physics, 3, 839–844.
MacInnis, I.N. and Brantley, S.L. (1992) The role of dislocations and surface mor-
phology in calcite dissolution. Geochimica et Cosmochimica Acta, 56, 1113– MANUSCRIPT RECEIVED AUGUST 17, 2001
1126. MANUSCRIPT ACCEPTED DECEMBER 24, 2001
Martinez, I., Zhang, J.Z., and Reeder, R.J. (1996) In situ X-ray diffraction of arago- MANUSCRIPT HANDLED BY MICHAEL E. FLEET

You might also like