Sie sind auf Seite 1von 20

Dolomite

It is probably safe to state that in 1982 no single model of dolomitization unequivocally


accounts for all aspects of any massively dolomitized ancient limestone. All models have
significant flaws, and our understanding of the dolomitization process and its relation to other
diagenetic processes (silicification, stylolitization, organic maturation, etc.) is imperfect.
Rather than advocate one solution over another, I will try to summarize some of the strengths
and weaknesses of several of the models which have been proposed.
As a starting point I will review several important aspects of dolomite mineralogy and
chemistry that place constraints on all models and that are sometimes overlooked.

Mineralogy

Dolomite is a rhombohedral carbonate with the ideal formula CaMg[C03)2 in which calcium
and magnesium occupy preferred sites. In the ideal mineral, planes of C03 anions alternate
with planes of cations with the c-axis of the crystal perpendicular to the alternating stacked
anion and cation planes. Ordering occurs by the additional alternation of cation planes
containing only calcium with cation planes containing only magnesium (Fig. 1). It is possible
to conceive of a mineral having the same composition as ideal dolomite ((Cao.6Mg0 6)C03) in
which all cation planes are alike, containing equal numbers of calcium and magnesium ions.
Such a mineral is not dolomite. Such a disordered arrangement of ions occupies more volume
than that of the ideal dolomite structure and is unstable with respect to an ordered phase.
Perhaps surprisingly, the two compounds just described, ideal dolomite and a disordered
1-to-l ratio Ca-Mg carbonate, are both rare in sedimentary rocks. Ideal dolomite rarely
comprises ancient dolomitic sediments and never modern sediments, and the completely
disordered polymorph does not occur at all. The dolomite which does occur in sedimentary
rocks is commonly Ca-rich, having compositions which range from about Ca(Cao.16Mgog4)(C03)2
to ideality, and/or exhibits weak, diffuse, X-ray diffraction, suggesting considerably less
structural order than its composition should dictate. With respect to ideal dolomite, all such
naturally occurring dolomite is metastable, and the capacity exists for reactions to occur
toward a more stable (more stoichiometric or better ordered) phase.
The term protodolomite was defined by Graf and Goldsmith (1956) as "single-phase
rhombohedral carbonates which deviate from the composition of the dolomite that is stable in
a given environment, or are imperfectly ordered, or both, but which would transform to
dolomite if equilibrium were established." Gaines (1977) modified the definition to include only
ordered phases. I recommended (1980) that the term be dropped altogether, since almost all
sedimentary dolomite is really protodolomite by Gaines' definition. What is important is not
what we call these natural materials, but what they really are.

1
J^ o
CARBONATE MAGNESIUM CALCIUM

Figure 1 Schematic representation of the crystal structure of dolomite showing the


alternation of cation and anion (carbonate) planes, and the alternation of calcium and
magnesium planes.

2
Hydrothermal experiments (Graf and Goldsmith, 1956; Goldsmith and Heard, 1961),
extrapolated to low temperature, demonstrate that calcite and dolomite are essentially ideal in
composition at 25 C (Fig. 2). In other words, any double carbonate crystal of Ca and Mg at
25 C which is not essentially pure dolomite is either metastable or unstable with respect to a
mixture of pure calcite plus pure dolomite. The same thing is true with respect to ideal
dolomite plus magnesite. The composition of phases which we observe at Earth's surface
define the range of metastability. Unstable phases are only observed as transient states in the
laboratory. In the case of dolomite, few phases containing more than about 8% excess calcium
(on a molar basis) have been reported to date, although the data are admittedly sparce.
Reeder (1981) has shown that the structure of various kinds of dolomite revealed by
transmission electron microscopy and electron diffraction can be classified into at least three
types. All structures are ordered, although the degree of order is variable and difficult to
quantify. The first, characteristic only of Holocene dolomite, consists of irregular "mosaics" on
a scale of tens or hundreds of Angstroms. The crystals are characterized by extremely high
densities of crystallographic faults and dislocations, and can be thought of as an aggregate of
"micro-crystals" whose compositions may vary, forming a very discontinuous lattice. This
leads to many unsatisfied or strained chemical bonds and to X-ray diffraction patterns with
broad, generally weak reflections. This kind of dolomite is also characterized by large trace
element substitutions, especially strontium (Behrens and Land, 1972), and sodium (Land and
Hoops, 1973). Qualitative data suggest that this material is extremely soluble compared to
better ordered forms of dolomite. My attempts to beneficiate samples composed of mixtures of
this kind of dolomite and aragonite (for example, supratidal crusts from Florida and the
Bahamas) by slow leaching in acetic acid resulted in only slight concentration of the dolomite
by selective solution of aragonite. C0 2 for isotopic analyses of Holocene dolomite is evolved
much faster than from finely ground ancient dolomite. All evidence suggests that Holocene
dolomite is a unique, highly soluble material. It is clearly a metastable phase, unknown (in an
unmodified form) in ancient rocks.
The second and most common kind of sedimentary dolomite exhibits a lamellar or "tweed"
structure when examined by transmission electron microscopy and electron diffraction, which
Reeder (1981) has interpreted as a structural and/or compositional modulation on a scale of
several hundred Angstroms (Fig. 3). At present this kind of dolomite is thought to consist of
two intimately intergrown lamellar domains parallel to the rhomb face with slightly different
structures and/or compositions. The texture resembles spinoidal decomposition, or solid state
unmixing on a scale of a few hundred angstroms from a single homogeneous precursor. The
exact structure and composition of the two domains or lamellae is not known, although one
must be more stable (and presumably more magnesium rich) than the other. This type of
dolomite is clearly metastable, but continued stabilization cannot proceed spontaneously
because it is limited by solid state diffusion. Continued stabilization can occur as a result of
solution-reprecipitation processes however, and it has been demonstrated that bulk Ca-rich
dolomites dissolve more rapidly than ideal dolomite (Busenberg and Plummer, 1982).
Continued stabilization toward a more stoichiometric dolomite would presumably be
promoted if pore fluids in the rock changed to enable dissolving out of the less stable, Ca-rich
domain. Porosity could easily increase under these conditions.

3
Ordered Dolomite

1000 -

Dolomite + Magnesite

800

600 -

TEMP.
(C) Lower limit of
400 - experimental data

Ranges of metastable phases


observed in nature
(

10 20 30 40 50 60

MOLE % MgC03

Figure 2 Stability relations in the system CaC03 MgC03


Figure 3 Dark field transmission electron micrograph of a calcian dolomite
(Caj i2Mg0 88(C02)2) of Eocene age. The prominent modulated structure is typical of sedimentary
dolomite, and such crystals are metastable with respect to ideal stoichiometric dolomite.
Photograph by Richard Reeder.

A third kind of dolomite is nearly ideal in composition, and when examined by transmission
electron microscopy and electron diffraction is observed to be homogeneous, consisting of
large single domains. This kind of dolomite is presently known mostly from ancient, deeply
buried sequences and from metamorphic rocks.
The philosophy that, like limestone, the diagenesis of dolomite is dominated by the
stabilization of metastable dolomitic phases, is relatively new. There is no question that
calcium-rich dolomite has the capacity to react to form crystals with a more stoichiometric
composition, but many important questions remain. What kinds of diagenetic environments
promote the reaction? Does stabilization to ideal dolomite take place all at once or in several
stages? How far from ideality must a phase be before it is prevented from further reaction for
kinetic reasons? Many of these questions must be answered both by laboratory work and by
careful mineralogical analysis of particular dolomites under investigation before thinking can
advance much further.

5
Aqueous solution equilibria

Several lines of evidence have been used to determine the solubility of dolomite at
sedimentary and early burial conditions. The data are complicated by the mineralogical
variations in dolomite already discussed. All metastable phases must be more soluble than
ideal dolomite, and variations in the degree of metastability can obviously occur.
Data have been derived from two sources, (1) high temperature experiments and (2) natural
dolomite aquifers. Of interest is the equilibrium constant, K, for reactions between the ideal
solids,

2CaC0 3 + Mg + + ^ CaMg(Co3)2 + Ca ++ K = (Ca ++ )/(Mg ++ )

or the calcium-to-magnesium activity ratio of a solution at equilibrium with calcite + dolomite


(as a function of temperature). Solutions more magnesium-rich than the equilibrium solution
should cause dolomitization of calcite, while solutions more calcium-rich should cause
dedolomitization.
Dolomite is easily synthesized hydrothermally at about 300 C, with reaction times of only a
few days. Progressively slower reaction is observed at lower temperatures and below about
100C very long experiments are required. Nobody has yet synthesized dolomite at
Earth-surface conditions (although a Dalmatian has!, Mansfield, 1980). Experimental data are
in reasonable agreement around 300C, and the molar Ca/Mg ratio of a solution in equilibrium
with calcite and dolomite is about 15. In other words, as temperature increases, dolomite
becomes increasingly less soluble than calcite. Any solution with a molar Ca/Mg ratio of less
than 15 is capable of dolomitizing at 300 C (Fig. 4)!
At lower temperatures, experimental data become more conflicting, the reason being, I
suspect, that metastable Ca-rich phases are much more easily formed. Kinetic experiments
(Land, 1967) have shown that the formation of a Ca-rich (metastable) dolomite rather than the
ideal phase is favored (within the stability field of dolomite) by (1) higher Ca/Mg ratio of the
solution, (2) lower solution concentration, and (3) lower temperature. Metastable Ca-rich
phases are more soluble and therefore will coexist with more magnesium-rich fluids (Helgeson
et al, 1978).
Hsu (1963), Holland et al, (1964), Barnes and Back (1964) and Langmuir (1971) all studied
the Ca/Mg ratio of natural dolomite aquifers, reasoning that equilibrium with dolomite would
eventually be reached as water recharged a dolomite aquifer and moved downdip at rates
typical for groundwater flow. Langmuir's compilation is plotted on Figure 4. The
extrapolation of Rosenburg and Holland's (1964) data to intercept Langmuir's low
temperature data is not too unreasonable if one accepts that the lower temperature
hydrothermal data points of Rosenburg and Holland may be displaced toward
magnesium-rich compositions because of formation of a non-ideal (more soluble) phase. The
reasonable agreement between low temperature and high temperature data ignores non-ideal
solution behavior, which is significant in the saline solutions Rosenburg and Holland used. But
at 300C, experiments at 2M, 1M and 0.5M solutions all yield similar results, suggesting the
effects are not large. Further support for extrapolation between the two types of data was

6
Temp. (C)
10
3.5 - Langmuir, 1971
25
CALCITE

50

3.0 -

DOLOMITE

100
2CaC03 + Mg++ ^ CaMg(C03)2 + Ca++
2.5 -

150
1000
T(K)

t \ 200

2.0 - Rosenberg and


Holland, 1964
250
Gaines (pers. comm.)
300
Land.
Rosenberg, Burt and Holland
1.5
-0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 logCa/Mg
I I I
5 10 25 Ca/Mg

Figure 4 Aqueous solution compositions presumed to be in equilibrium with calcite plus


dolomite.
obtained by Pakhomov and Kisson (1973) (reproduced in Carpenter, 1980) who plotted the
Ca/Mg ratio of saline formation water from the Russian platform versus temperature. Despite
the fact that they totally ignored rock composition (calcite plus dolomite may not both have
been present to control the solution composition), and obtained considerable scatter, their
regression line essentially connects Rosenburg and Holland's and Langmuir's data! Until
further experimental work is conducted (which must include characterization of the dolomite
phase) the data presented in Figure 4 are all that are available. They are consistent both with a
gross oversaturation of seawater with dolomite, and the Mg-depleted nature of most saline
formation water.
The reason for the gross oversaturation of seawater with respect to dolomite ultimately lies
in the kinetic problem of nucleating and growing the ordered crystal (Goldsmith, 1953). The
molar Ca/Mg ratio of seawater (0.19) is apparently incapable of causing dolomitization at
observable rates. By either decreasing the molar Ca/Mg ratio of seawater (say by gypsum
precipitation) or decreasing the activity Ca/Mg ratio the kinetic constraints can be overcome,
at least to the point of being able to nucleate and grow a poorly crystalline Ca-rich phase. The
activity Ca/Mg ratio of seawater is 0.18 (Berner, 1971), and can be decreased by dehydrating
the Mg+ + ion (Usdowski, 1968) or by removing components which form strong ion pairs with
Mg+ + (for example, S0 4 = , Baker and Kastner, 1981). These factors do not alter the equilibrium
relations (Fig. 4) and only provide the kinetic "push" to form the initial phase. The
early-formed phase can then stabilize by further reaction.
Another variable in the dolomitization process which needs additional confirmation is the
role of organic material, particularly dissolved organic acids. Dissolved organic acids are
known to control the kind of calcium carbonate which precipitates from solution. Increased
organic acid content favors Mg-calcite over aragonite precipitation (Kitano and Kanamori,
1966). Although algal processes have been invoked as being able to cause dolomitization
(Gebelein, 1973), the "organic gremlin" is neither proven nor disproven.

Stable Isotopic Geochemistry

Most current evidence supports the contention that sedimentary dolomite is enriched in 180
about 3 to 4 ppt with respect to a co-existing calcite in the range of sedimentary and burial
diagenetic temperatures of normal interest (Land, 1980). Little evidence exists for dolomite
replacement of calcite without change of isotopic composition (Katz and Matthews, 1977). The
fact that many ancient dolomites are significantly depleted in 180 is best explained by
stabilization of an earlier-formed phase during burial (Fig. 5). The isotopic composition of the
dolomite comprising sedimentary rocks is controlled both by the chemistry of the latest
recrystallization (stabilization) event and by the chemistry of the precursor (aragonite,
Mg-calcite, calcite and/or dolomite). Dolomite rarely recrystallizes homogeneously in an open
aqueous chemical system, accurately recording the conditions of recrystallization, just as it
almost never accurately retains the chemistry of the precursor. Recrystallization may be
incomplete, leaving an inhomogeneous rock, and the composition of the replaced phase may
"contaminate" the replacing phase (Land, 1980). The practical problem of analyzing intimate
mixtures of dolomite of slightly different compositions is not yet solved.

8
200

160-

Temp. 12 o
(C)

80

40-

<5180 Dolomite (PDB scale)

Figure 5 Oxygen isotopic composition of dolomite as a function of temperature and 5180


water (curved lines). The histogram at top was constructed from published analyses (Land,
1980) and shows the wide range of conditions under which dolomite formed and/or stabilized.

9
Trace Element Geochemistry

Trace element partitioning is complicated by kinetic factors. The ratio of the concentration
of a trace element in a crystal to the concentration of the element for which it substitutes (say
Sr/Ca) is dependent on the concentration ratios of the elements in the solution from which the
crystal forms, on temperature, on pressure (usually ignored), and on other variables such as
the rate of crystal growth. The distribution coefficient "D" in the following equation is thus a
function of variables which are not always easy to define:

(Sr/Ca)crygtal = Dx(Sr/Ca)
solution

Modern marine and hypersaline dolomite has an Sr content of about 600 ppm (Behrens and
Land, 1972), yet few ancient dolomites contain more than 200 ppm Sr, even when presumed to
be initially of hypersaline origin. Although it was once assumed that removal of the trace
elements by flushing with a low Sr (meteoric) water was required (Land, 1973), this is no longer
acceptable for all ancient dolomite.
As an example of this problem, Bein and Land (1982) studied Permian San Andres dolomite
from the subsurface in north Texas, where dolomite beds are intimately interbedded with
bedded halite and anhydrite. Both halite and anhydrite display sedimentary structures
indicating a primary subaqueous origin, and both contain trace elements (Br in halite and Sr in
anhydrite) indicative of primary precipitation. It seems clear that thin dolomite beds
intimately interbedded with and "entombed" by primary evaporites could never have formed
from or been modified by low Sr (meteoric) water. Yet the dolomites all contain less than 200
ppm Sr. Bein and Land suggest that although the original dolomite may have resembled
Holocene analogs (about 600 ppm Sr), during burial it stabilized to a more ordered structure,
expelling Sr to form celestite. In other words, at least two distribution coefficients apply to
this situation, one for the formation of the original phase, and a second (lower) for the
stabilization reaction to a more ordered, stoichiometric phase.
Because of these kinetic problems which plague other sedimentary phases as
well anhydrite (Kushnir, 1980), halite (Holser, 1979), trace element analyses of dolomite are
of limited practical value today. Hopefully, more experimental work will rectify this situation.

Mechanisms of Dolomitization

Clearcut petrographic evidence indicates that most dolomite initially forms by replacing a
precursor carbonate. That is, a fluid simultaneously imports Mg++, dissolves the precursor
phase, precipitates dolomite, and exports Ca++. Of course, the situation is actually more
complex due to the import and export of other components such as other trace elements and
their isotopes (for example, 87Sr/86Sr), carbon and oxygen isotopes, C02, etc. Because of
considerable compositional differences between dolomite and any presumed precursor (calcite,
aragonite, or Mg-calcite), considerable fluid transport is required. Advection (fluid flow) must
accomplish most of the transport, although diffusion may play an important part on a local
scale. Models for dolomitization are therefore basically hydrologic models. Before discussing

10
^ ^r seawater

STORM RECHARGE

EVAPORATIVE DRAWDOWN

Figure 6 Only two models of sabkha hydrology can apply if seawater is the hydrologic
baselevel, and simultaneously the source for magnesium. Elevation of seawater onto the
sabkha surface by storms, or by lowering the baselevel, provides the elevation head to move
water back to the sea (reflux). Evaporative drawdown can only be a transient condition as the
depression is rapidly filled with salts.

these, we should not completely ignore' 'primary" dolomite.


Replacement is a dissolution-precipitation process, albeit on a sub-microscopic scale. The
dolomite which replaces a precursor is a precipitate in every sense of the word, and is primary
in the sense that it was not there before. It is not primary only in the sense that it occupies
space previously occupied by another solid phase. "Primary dolomite" is usually defined as
crystals which nucleate from solution and either accumulate as primary sediment or
precipitate into megascopic pores as cement, and so displace only fluid as they grow.
Dolomitic rocks which contain pore-filling dolomite cements do, in every sense of the word,
consist in part of primary dolomite. Dolomitic rocks are often very inhomogeneous as a result
of containing several generations of dolomite.
Primary dolomite micrite may not be as uncommon as current thinking presumes. Baffin
Bay, Texas, provides a possible example. Essentially pure, laterally extensive dolomite beds
up to 4 cm thick occur in early Holocene subtidal laminated terrigenous muds. Cores from the

11
upper part of the Bay sequence, from the sea floor to about 7 m below, are extremely well
laminated, documenting alternating periods of aragonite (and rarely Mg-calcite) precipitation
and terrigenous deposition, which occurred during and after storms (often hurricanes). The
bay is normally hypersaline except after hurricanes, and so the deposition of chemical
precipitates during hypersaline periods and the deposition of terrigenous material
accompanying runoff accounts for the laminations, and the hypersalinity for their
preservation. The middle part of the sequence, from about 7 to 13 m below sea floor, formed in
about 5 m of water about 3500 -1000 years ago (Behrens, 1974), and is texturally similar
except for the presence of dolomite beds. Very little terrigenous material is present within the
dolomite beds, ruling out any kind of a mixing model since fresh water would have contributed
terrigeneous mud. Interstitial water analyses of the very impermeable sediments, obtained by
hydraulic squeezer, have a relatively uniform chlorinity (36 ppt), molar Ca/Mg ratio (0.15) and
<5180 (+ 2 ppt), very similar to Baffin Bay water during normal summers. Only small amounts
of interstitial gypsum are present in the sediment and no beds of gypsum occur. Clearly a 4
cm-thick dolomicrite could not have formed after burial because the pore water shows no
strong depletion in magnesium. The beds must have formed at the sea floor, either by primary
precipitation or by complete replacement of some precursor (Mg-calcite?), prior to being buried
by terrigeneous influx associated with storms. Although it is probable that the early Baffin
Bay was silled, and sulfate reduction in the (stratified?) bottom water may have been
important in reducing the Ca/Mg activity ratio of the water (Baker and Kastner, 1980), the
possibility that the dolomite is a primary precipitate cannot be excluded. The possible
importance of primary dolomicrite should not be ignored.
Three hydrologic models for dolomitization are currently "in vogue." All require a potential
field to move fluid through the rocks and an "inexhaustible" source of magnesium. The
following calculation serves to illustrate the magnitude of fluid flow and magnesium required.
Assume a typical carbonate sediment is to be dolomitized. A typical sediment contains
about 6.3 mole percent MgC03 (Land, 1973, Table 2), and has about 40% porosity. As a place to
start, assume that seawater is to be the dolomitizing agent. A cubic meter therefore contains:
a) 400 liters of seawater x 1.025 Kg water/liter water = 4.1 x 102 kg seawater; and
b) 600,000 cu cm of limestone consisting of:

570,885 cu cm CaC03 or 1.545 x 104 moles CaC03 (36.94 cu cm/mole)


29,115 cu cm MgC03 or 1.039 x 103 moles MgC03 (28.02 cu cm/mole).

4.1 x 102 kg of seawater contains 4.26 moles of Ca+ + and 22.1 moles of Mg++. If the sediment
reacts with the water to reach equilibrium (calcite -I- dolomite + a solution having a Ca/Mg ~
1, Fig. 4), then the interstitial water will provide 8.92 additional moles of magnesium and the
rock will contain 6.7% dolomite of ideal composition. 99% of the dolomite is derived from the
magnesium originally in the Mg-calcites and slightly less than 1 % is derived from the
magnesium in the interstitial seawater. To completely dolomitize the remaining CaC03,7.19 x
103 moles of Mg+ + must be added. E ach pore volume of "new'' seawater can provide 8.92 moles
of Mg+ + for dolomitization (the water can only provide magnesium until the Ca/Mg ratio is
increased to about 1Fig. 4at which point it reaches equilibrium with calcite + dolomite).

12
Therefore, 807 pore volumes of seawater are required to completely dolomitize 1 cu m of
sediment. If seawater diluted 10 times with meteoric water (say in a mixing zone) is utilized,
then 8.1 x 103 pore volumes are needed. If seawater having an Mg content of about 8 x 10"1 and
a Ca content of about 8 x 10 "moles/Kg is utilized (a typical brine which has precipitated
gypsum and evaporated to the point of halite saturation), then only 44 pore volumes are
needed. If the various solutions do not reach equilibrium with calcite + dolomite (the Mg/Ca
ratio does not fall to 1), or the brine has not reached halite saturation, then proportionately
more pore volumes of fluid are required. No porosity reduction has been achieved, and if
dolomite cementation occurs, additional fluid flow is required.

Reflux

Reflux, as defined by Adams and Rhodes (1960) occurs when "hypersaline brines eventually
become heavy enough to displace the connate waters and seep slowly downward through the
slightly permeable carbonates at the lagoon floor." Examination of Holocene sabkhas has
suggested that downward moving water driven solely by potential energy resulting from
increased density of the fluid at constant head is probably not as important as the increased
head caused by elevation of water onto the sabkha surface by storms.
Hsu and Siegenthaler (1964) summarized various ideas of sabkha hydrology. Basically,
considering a sabkha which extends relatively far along strike relative to its width (a two
dimensional system), the directions of water movement are quite limited. At any point in the
sabkha, water can either move up or down, seaward or landward (Fig. 6). It is assumed that an
infinite reservoir of magnesium (seawater) is available at some constant level at the margin of
the sabkha. Only two processes can move seawater (the source of magnesium) landward in the
absence of interaction with an independent underlying aquifer system, namely storm recharge
and evaporative drawdown. Evaporative drawdown, or the lowering of the water table by
evaporation, can only occur if the landward part of the sabkha is depressed below sea level by
subsidence, compaction and/or wind deflation. Unevaporated seawater must be kept from
flooding the depression by some sort of sill, either a physical barrier or a long distance. In any
case, landward flow of seawater into a depression will result in rapid evaporation and
consequent filling of the basin by evaporite minerals, effectively halting flow by eliminating
the head difference. The amount of water required to produce 1 cu m of gypsum is about
sufficient to completely dolomitize 1 cu m of carbonate sediment. Evaporative drawdown
(possibly aided by capillary withdrawal) is, at best, a transient condition and is self-limiting.
Storm recharge, however, can continuously (geologically speaking) drive water up onto the
sabkha, where it evaporates and flows seaward, driven by elevation head, and aided by its
increased density. Such a mechanism dominates modern sabkhas (McKenzie, Hsu, and
Schneider, 1980; Amdurer and Land, 1982). In the case of the Trucial Coast of the Persian
Gulf, dolomitization takes place only in the storm recharge zone, and the amount of dolomite
correlates with the frequency of recharge (Patterson and Kinsman, 1982).
Considerable amounts of gypsum may be precipitated as the result of brine evolution. For
example, using the figures previously discussed, 44 pore volumes of halite-saturated brine
were required to dolomitize 1 cu m of sediment. About 1 cu m of gypsum would have

13
DOLOMITE

200

SUPERSATURATED
100
UNDERSATURATED

50

ZONE OF DOLOMITIZATION

I 1 j -
0 20 40 60 80 100
percent seawater

Figure 7 Percent saturation for mixtures of seawater and a typical meteoric groundwater
having a P = 102 atmospheres (after Plummer, 1975).

precipitated from the volume of seawater required to generate that much brine, leading to a
gypsum-to-dolomite volume ratio of one.
Advantages of the reflux mechanism are the rapidity with which dolomite can be formed as
documented by Holocene studies, and the relatively smaller volumes of water required due to
its magnesium-rich nature (Sears and Lucia, 1980). This mechanism clearly dominates in
evaporitic settings. In the absence of evaporites the model is more constrained, barring
fluctuations of the Ca/Mg ratio and/or the sulfate content of seawater. The efficient removal of
calcium by the formation of surficial algal micrite prior to evaporative concentration can also
suppress CaSO< precipitation (Amdurer and Land, 1982). In addition, it is not at all clear that
reflux can operate on the regional scale for which it was first proposed. The small fluid

14
potentials caused solely by density differences apparently cannot move water very far through
sediments of relatively low permeability. Elevation head is required, and in addition to storm
recharge it might easily be accomplished by periodic lowering of the reservoir of seawater
either by a local mechanism (say evaporation of a restricted sea) or on a larger scale
(eustatic/tectonic), draining of the sabkhas periodically in the same way modern coastal plains
were drained during Pleistocene glacial events.

Meteoric Mixing

In order to account for evaporite-free dolomite sequences, the mixing of meteoric water
(providing the driving force through elevation head) with seawater (providing the magnesium)
has been advocated (Hanshaw, Back, and Deike, 1971; Land, 1973). Geochemical
considerations (Fig. 7) (Plummer, 1975) suggest that the mechanism is plausible even though
much longer times are required for dolomitization (Sears and Lucia, 1980). Although examples
of Holocene mixing-zone dolomite (mostly as cements!) continue to be found (Magaritz et al,
1980), a major problem with the model is explaining why dolomite is not more common, since
mixing of seawater and meteoric water is a ubiquitous worldwide process. The model appar-
ently requires a relatively stable hydrologic setting to establish sufficient continuous recharge
for establishment of a mixing cell with seawater over a long period of time to drive the dolomit-
ization reaction. Kinetic problems are overcome by reducing the Ca/Mg activity ratio of the
mixture through lowering of the ionic strength. This may not be too much of a problem in a
subtropical setting as, say, tidal flats prograde across a shelf leaving behind vast areas for
recharge. But in an arid climate the model is difficult to apply unless large adjacent coastal
plains provide the recharge zone and evaporites are sealed off from the actively circulating
water.
The evaporative concentration of continental water accounts for playa-type dolomite includ-
ing the Coorong examples (von der Borch, Lock and Schwebel, 1975).

Burial Diagenesis

Dolomite can clearly form as a directly precipitated late cement, as exemplified by studies of
sandstone burial diagenesis (Boles, 1978; Land and Dutton, 1978). The dolomite is commonly
ferroan, and can approach ankerite in composition, reflecting the large amount of ferrous iron
commonly present in the terrigenous system. Although it is true that shales in a sedimentary
basin are possible sources for nearly every conceivable component required for any conceivable
kind of diagenesis, it is not clear that they are sources for magnesium. In fact, the
precipitation of chlorite within the shales may be a local sink for magnesium. Saline formation
waters are typically very magnesium-poor, and on the whole commonly approach
calcite-dolomite equilibrium (Pakhomov and Kissin, 1973). Supplying large amounts of
magnesium from a water nearly in equilibrium with calcite plus dolomite requires vast
amounts of water, a definite problem, especially in relatively impermeable rocks. In addition,
Figure 4 indicates that if a water initially in equilibrium with calcite plus dolomite moves

15
updip (and cools), it becomes undersaturated with dolomite and will either dissolve dolomite or
dedolomitize. This exact subsurface reaction has been observed by Land and Prezbindowski
(1981) and Budai (1981).
Therefore, at the present time, the formation of large amounts of new replacement dolomite
is difficult by this mechanism. No large-scale source for magnesium has been identified.
Moving magnesium around within a basin without producing any net new dolomite appears to
be quite possible, but in this case the "new" replacement dolomite or cement must be balanced
by either "new" dedolomite or by secondary porosity somewhere within the basin. The
mobility of calcium, magnesium and dissolved carbonate after burial must not be disregarded.
Shales are rapidly "decalcified" during burial (Hower et al, 1976) and provide a large-scale
source for new carbonate phases. But since calcium loss exceeds magnesium loss by at least a
factor of 6, much more calcite than dolomite is involved in the process. Sandstone diagenesis
can involve immense quantities of carbonate which is both precipitated and removed (to form
secondary porosity). Sandstones can be carbonate-cemented, decemented and then recemented
(Milliken et al, 1982), and carbonates probably undergo similar complex histories. Late
secondary porosity development in carbonates is known (Moore and Druckman, 1981), and
some textures in deeply buried carbonates may be the result of selective dissolution of calcite,
leaving the dolomitic component of the rocks as an "insoluble residue" (Wanless, 1979).
It is important to "decouple" the process of dolomitization/dedolomitization (controlled by
the Ca/Mg of the solution) from cementation/secondary porosity generation (controlled by the
acidity of the solution). The dolomitization process is rarely C03=-conservative (Weyl, 1960;
Degens and Epstein, 1964). A solution with a low Ca/Mg ratio and capable of dolomitizing can
either cause net cementation or net solution, depending on changes occurring in the total
dissolved carbonate content of the solution as it moves through the rocks. Addition of C0 2 by
organic maturation can cause net solution, whereas loss of C0 2 to adjacent strata of lower
carbonate content can cause net precipitation. Thus dolomitization can either result in
porosity decrease (by cementation and/or by compaction accompanying recrystallization), or
porosity increase (secondary porosity formation). The same is true of the dedolomitization
reaction.
Classic dolomite reservoirs containing intercrystalline porosity may possibly result from
recrystallization of a metastable Ca-rich precursor phase induced by a C02-rich (corrosive)
solution. Some or all of the more Ca-rich (more soluble) domains of the metastable phase may
be lost to the solution, and additional dolomite may even be dissolved. The less soluble
component must recrystallize, and intercrystalline porosity results from the volume loss of the
Ca-rich domains. It is possible that such situations may even be "self-reservoiring" in the
sense that C0 2 evolved during early maturation may be responsible for creating the reservoir
by dolomite recrystallization!

Other Possibilities

We should be careful about being too actualistic in our approach to dolomitization. Only 25
years ago, we thought that essentially no Holocene dolomite existed (Fairbridge, 1957). Each
case of Holocene dolomitization has resulted in considerable over-reaction and

16
"bandwagon-jumping" soon after the discovery.
One intriguing possibility, which is gaining considerable support recently, is that "the
present is a lousy key to the past because seawater has changed." The observation that the
percentage of dolomite in carbonate rocks increases as we go back in geologic time was
originally attributed to more time available for dolomitization (the source of magnesium was
not specified) (Chilingar, 1956). Changes in the composition of seawater resulting in times of
"easier" dolomitization in the past cannot be discounted. Tucker (1982) recently suggested a
primary origin for a Pre-Cambrian oosparite (oolitic grainstone) composed entirely of dolomite
(including the "spar"!). Changes in salinity, in Ca/Mg ratio, SC%= concentration and PC02 have
all been invoked (Sandburg, 1975; Baker and Kastner, 1981; Mackenzie and Pigott, 1981), and
sympathetic variation of several components may be particularly effective, and ultimately
related to crustal cycles.

Conclusions

No panaceas exist for dolomitization. Each case must be studied on its own merits, and
many scenarios exist. Modern scenarios begin to break down if seawater and/or sediment
compositions have evolved with time. Reflux can account for the initial formation of many
evaporite-related dolomites but since the poorly ordered phases formed in hypersaline
environments are not found in ancient rocks, recrystallization must occur. Mixing zone
dolomitization is capable of upgrading early hypersaline phases to a more stable phase, but is
not necessary as "isochemical" recrystallization can occur in saline brines. Mixing zones are
capable of producing dolomite cements and new replacement phases, given enough time and
with sufficient recharge zones. Burial diagenesis can generate dolomite cements (commonly
ferroan), induce recrystallization of previously formed, metastable phases, and move
previously formed dolomite from place to place. Recrystallization can take place in essentially
closed chemical systems, or in partly open systems resulting in gross changes in the chemistry
of the dolomite and in the selective removal of either calcite or dolomite from the sequence.
Few (if any) carbonate rocks, dolomitized or not, exist as they were originally deposited.
Most have resulted from one or more processes of formation, and at least one stabilization
(recrystallization) event.

Acknowledgements

Several students and colleagues critiqued earlier versions of the manuscript and offered
valuable corrections, including James Anderson, David Budd, Bob Folk, Donald Miser, and
Richard Reeder. Richard Reeder kindly provided Figure 3. Support of the Geology Foundation
of the University of Texas at Austin is gratefully acknowledged.

17
REFERENCES
Adams, J. E. and M. L. Rhodes, 1960, Dolomitization by seepage refluxion: AAPG Bull., v. 44,
p. 1912-1920.
Amdurer, M. and L. S. Land, 1982, Geochemistry, hydrology and mineralogy of the sand bulge
area, Laguna Madre flats, South Texas: Jour. Sed. Petrology, v. p.
Baker, P. A. and M. Kastner, 1981, Constraints on the formation of sedimentary dolomite:
Science, v. 213, p. 214-216.
Barnes, I. and W. Back, 1964, Dolomite solubility in groundwater: U. S. Geol. Survey Prof.
Paper 475-D, p. 179-180.
Behrens, E. W., 1974, Holocene sea level rise effect on the development of an estuarine
carbonate depositional environment: Memoires de 1'Justitat de Geologie du Bassin
d'Aquitaine, No. 7, p. 337-341.
and L. S. Land, 1972, Subtidal Holocene dolomite, Baffin Bay, Texas: Jour. Sed.
Petrology, v. 42, p. 155-161.
Bein, A. and L. S. Land, 1982, San Andres carbonates in the Texas panhandle; sedimentation
and diagenesis associated with magnesium-calcium-chloride brines: Austin, Texas, Univ. of
Texas, Bureau of Econ. Geology, Rept. of Invest., No. 121,48 p.
Berner, R. A., 1971, Principles of chemical sedimentology: McGraw-Hill, 240 p.
Boles, J. R., 1978, Active ankerite cementation in the subsurface Eocene of southwest Texas:
Contrib. Mineralogy and Petrology, v. 68, p. 13-22.
Budai, J. M., 1981, Subsurface dedolomitization of the Madison limestone, Wyoming: Geol.
Soc. America Abs. with Programs, p. 419.
Busenberg, E. and L. N. Plummer, 1982, The kinetics of dissolution of dolomite in CO2-H20
systems at 1.5 to 65C and 0 to 1 atm Pco2: Am. Jour. Sci., v. 282, p. 45-78.
Carpenter, A. B., 1980, The chemistry of dolomite formation I; the stability of dolomite, in D.
H. Zenger, J. B. Dunham, and R. L. Ethington, eds., Concepts and models of dolomitization:
SEPM Spec. Pub. No. 28, p. 111-121.
Chilingar, G. V., 1956, Relationship between Ca/Mg ratio and geologic age: AAPG Bull., v. 40,
p. 2256-2266.
Degens, E. T, and S. Epstein, 1964, Oxygen and carbon isotope ratios in coexisting calcites
and dolomites from recent and ancient sediments: Geochim. et Cosmochim. Acta, v. 28, p.
23-44.
Fairbridge, R. W., 1957, The dolomite question, in R. J. LeBlanc and J. G. Breeding, eds.,
Regional aspects of carbonate deposition: SEPM Spec. Pub. No. 5, p. 125-178.
Gaines, A. M., 1977, Protodolomite redefined: Jour. Sed. Petrology, v. 47, p. 543-546.
, 1978, Reply, protodolomite redefined: Jour. Sed. Petrology, v. 48, p. 1009-1011.
Gebelein, C. D., 1973, Algal origin of dolomite laminations in stromatolitic limestone: Jour.
Sed. Petrology, v. 43, p. 603-613.
Goldsmith, Jr., 1953, A "simplexity principle" and its relation to "ease" of crystallization:
Jour. Geology, v. 62, p. 439-451.
and H. C. Heard, 1961, Subsolidus phase relations in the system CaC03-MgC03: Jour.
Geology, v. 69, p. 45-74.

18
Graf, D. L. and J. R. Goldsmith, 1956, Some hydrothermal syntheses of dolomite and
protodolomite: Jour. Geology, v. 64, p. 173-186.
Hanshaw, B. B., W. Back, and R. G. Deike, 1971, A geochemical hypothesis for dolomitization
by groundwater: Econ. Geology, v. 66, p. 710-724.
Helgeson, H. C , et al, 1978, Summary and critique of the thermodynamic properties of
rock-forming minerals: Am. Jour. Science, v. 278-A, 229 p.
Holland, H. D., et al, 1964, On some aspects of the chemical evolution of cave waters: Jour.
Geology, v. 72, p. 36-67.
Holser, W. T., 1979, Trace elements and isotopes in evaporites, in R. G. Burns, ed., Marine
minerals: Mineralog. Soc. of America Short Course Notes, v. 6, p. 295-346.
Hower, R., et al, 1976, Mechanism of burial metamorphism of argillaceous sediment; 1
Mineralogical and chemical evidence: Geol. Soc. America Bull., v. 87, p. 725-737.
Hsu, K. J., 1963, Solubility of dolomite and composition of Florida groundwaters: Jour.
Hydrology, v. 1, p. 288-310.
and C. Siegenthaler, 1969, Preliminary experiments on hydrodynamic movement
induced by evaporation and their bearing on the dolomite problem: Sedimentology, v. 12, p.
11-26.
Katz, A. and A. Matthews, 1977, The dolomitization of CaC03; an experimental study at
252-295C: Geochim. et Cosmochim. Acta, v. 41, p. 297-308.
Kitano, Y, and N. Kanamori, 1966, Synthesis of magnesian calcite at low temperatures and
pressures: Geochem. Jour., v. 1, p. 1-10.
Kushnir, J., 1980, The coprecipitation of strontium, magnesium, sodium, potassium and
chloride ions with gypsum; an experimental study: Geochim. et Cosmochim. Acta, v. 44, p.
1471-1482.
Land, L. S., 1967, Diagenesis of skeletal carbonates: Jour. Sed. Petrology, v. 37, p. 914-930.
, 1973, Contemporaneous dolomitization of Middle Pleistocene reefs by meteoric water,
north Jamaica: Bull. Marine Science, v. 23, p. 64-92.
, 1980, The isotopic and trace element geochemistry of dolomite; the state of the art, in
D. H. Zenger, J. B. Dunham, and R. L. Ethington, eds., Concepts and models of
dolomitization: SEPM Spec. Pub. No. 28, p. 87-110.
and G. K. Hoops, 1973, Sodium in carbonate sediments and rocks; a possible index to
the salinity of diagenetic solutions: Jour. Sed. Petrology, v. 43, p. 614-617.
and S. P. Dutton, 1978, Cementation of a Pennsylvanian deltaic sandstone; isotopic
data: Jour. Sed. Petrology, v. 48, p. 1167-1176.
and D. R. Prezbindowski, 1981, The origin and evolution of saline formation water,
Lower Cretaceous, south-central Texas, U. S. A.: Jour. Hydrology, v. 54, p. 54-71.
Langmuir, D. L., 1971, The geochemistry of some carbonate ground waters in central
Pennsylvania: Geochim. et Cosmochim. Acta, v. 35, p. 1023-1045.
Mackenzie, E T. and J. D. Pigott, 1981, Tectonic controls of Phanerozoic sedimentary rock
cycling: Jour. Geol. Soc. London, v. 138, p. 183-196.
Mansfield, C. G., 1980, A urolith of biogenic dolomite - another clue in the dolomite mystery:
Geochim. et. Cosmochim Acta., v. 44, p. 829-840.
McKenzie, J. A., K. J. Hsu, and J. E Schneider, 1980, Movement of subsurface waters under

19
the sabkha, Abu Dhabi, U. A. E., and its relationship to evaporative dolomite genesis, in D.
H. Zenger, J. B. Dunham, and R. L. Ethington, eds., Concepts and models of dolomitization:
SEPM Spec. Pub. No. 28, pp. 11-30.
Magaritz, M., et al, 1980, Dolomite formation in the seawater-freshwater interface: Nature, v.
287, p. 622-624.
Milliken, K. L., L. S. Land, and R. G. Loucks, 1981, History of burial diagenesis determined
from isotopic geochemistry, Frio Formation, Brazoria County, Texas: AAPG Bull., v. 65, p.
1397-1413.
Moore, C. H. and K. Druckman, 1981, Burial diagenesis and porosity evolution, Upper
Jurassic Smackover, Arkansas and Louisiana: AAPG Bull., v. 65, p. 597-628.
Pakhomov, S. I., and I. G. Kissin, 1973, Hydrogeochemistry of magnesium in deep aquifer
zones: Akad. Nauk SSSR Doklady, v. 209, p. 205-208. (English translation, 1974, American
Geological Institute).
Patterson, R. J. and D. J. J. Kinsman, 1982, Formation of diagenetic dolomite in coastal
sabkha along the Arabian (Persian) Gulf: AAPG Bull., v. 66, p. 28-43.
Plummer, L. N., 1975, Mixing of seawater with calcium carbonate ground water: Geol. Soc. of
America Mem. 142, p. 219-236.
Reeder, R. J., 1981, Electron optical investigation of sedimentary dolomites: Contr.
Mineralogy Petrology, v. 76. p. 148-157.
Rosenberg, P. E., D. M. Burt, and H. D. Holland, 1967, Calcite-dolomite-magnesite stability
relations in solutions; the effect of ionic strength: Geochim. et Cosmochim. Acta, v. 31, p.
391-396.
Rosenburg, P. E., and H. D. Holland, 1964, Calcite-dolomite-magnesite stability relations in
solutions at elevated temperatures: Science, v. 145, p. 700-701.
Sandburg, E A., 1975, New interpretations of Great Salt Lake ooids and of ancient
non-skeletal carbonate mineralogy: Sedimentology, v. 22, p. 497-537.
Sears, S. 0., and F. J. Lucia, 1980, Dolomitization of northern Michigan Niagara reefs by brine
refluxion and freshwater/seawater mixing, in D. H. Zenger, J. D. Dunham, and R. L.
Ethington, eds., Concepts and models of dolomitization: SEPM Spec. Pub. No. 28, p.
215-235.
Tucker, M. E., 1982, Precambrian dolomites; petrographic and isotopic evidence that they
differ from Phanerozoic dolomites: Geology, v. 10, p. 7-12.
Usdowski, H. E., 1968, The formation of dolomite in sediments, in G. Muller, and G. M.
Friedmann, eds., Recent developments in carbonate sedimentology in central Europe:
Springer-Verlag, p. 21-32.
von der Borch, C. C , D. E. Lock, and D. Schwebel, 1975, Groundwater formation of dolomite
in the Coorong region of South Australia: Geology, v. 3, p. 283-285.
Wanless, H. R., 1979, Limestone response to stress; pressure solution and dolomitization: Jour.
Sed. Petrology, v. 49, p. 437-462.
Weyl, K., 1960, Porosity through dolomitization; conservation-of-mass requirements: Jour.
Sed. Petrology, v. 30, p. 85-90.

20

Das könnte Ihnen auch gefallen